question
stringlengths
15
150
question_id
int64
258k
2.12M
score
int64
1
208
link
stringlengths
60
131
body
stringlengths
101
9.72k
answers
listlengths
1
288
verbolised
stringlengths
650
683k
Axiom of choice, Banach-Tarski and reality
260,057
65
https://mathoverflow.net/questions/260057/axiom-of-choice-banach-tarski-and-reality
<p>The following is not a proper mathematical question but more of a metamathematical one. I hope it is nonetheless appropriate for this site.</p> <p>One of the non-obvious consequences of the axiom of choice is the <a href="https://en.wikipedia.org/wiki/Banach%E2%80%93Tarski_paradox">Banach-Tarski paradox</a> and thus the existence of non-measurable sets.</p> <p>On the other hand, there seem to be models of Zermelo-Fraenkel set theory without axiom of choice where every set would be measurable.</p> <p>What does this say about the "plausibility" of the axiom of choice? Are there reasons why it is plausible (for physicists, philosophers, mathematicians) to believe that not all sets should be measurable? Is the Banach-Tarski paradox one more reason why one should "believe" in the axiom of choice, or is it on the opposite shedding doubt on it?</p>
[ { "answer_id": 260072, "score": 120, "text": "<p>There are <em>two</em> ingredients in the Banach-Tarski decomposition theorem:</p>\n<ol>\n<li>The notion of <em>space</em>, together with derived notions of <em>part</em> and <em>decomposition</em>.</li>\n<li>The axiom of choice.</li>\n</ol>\n<p>Most discussion about the theorem revolve around the axiom of choice. I would like to point out that the notion of space can be put under scrutiny as well.</p>\n<p>The Banach-Tarski decomposition of the sphere produces non-measurable parts of the sphere. If we <em>restrict</em> the notion of &quot;part&quot; to &quot;measurable subset&quot; the theorem disappears. For instance, if we move over into a model of set theory (without choice) in which all sets are measurable, we will have no Banach-Tarski. This is all well known.</p>\n<p>Somewhat amazingly, we can make the Banach-Tarski decomposition go away by <em>extending</em> the notion of subspace, and keep choice too. Alex Simpson in <a href=\"https://web.archive.org/web/20160909042100/http://homepages.inf.ed.ac.uk/als/Research/Sources/mrs.pdf\" rel=\"noreferrer\">Measure, Randomness and Sublocales</a> (Annals of Pure and Applied Logic, 163(11), pp. 1642-1659, 2012) shows that this is achieved by generalizing the notion of topological space to that of <a href=\"https://ncatlab.org/nlab/show/locale\" rel=\"noreferrer\">locale</a>. He explains it thus:</p>\n<blockquote>\n<p>&quot;The different pieces in the partitions defined by Vitali and by Banach and Tarski are deeply intertangled with each other. According to our notion of “part”, two such intertangled pieces are not disjoint from each other, so additivity does not apply. An intuitive explanation for the failure of disjointness is that, although two such pieces share no point in common, they nevertheless overlap on the topological “glue” that bonds neighbouring points in <span class=\"math-container\">$\\mathbb{R}^n$</span> together.&quot;</p>\n</blockquote>\n<p>Peter Johnstone explained in <a href=\"http://pointlesstopology.com/the-point-of-pointless-topology.pdf\" rel=\"noreferrer\">The point of pointless topology</a> why locales have mathematical significance that goes far beyond fixing a strange theorem about decomposition of the sphere. Why isn't everyone using locales? I do not know, I think it is purely a historic accident. At some point in the 20th century mathematicians collectively lost the idea that there is more to space than just its points.</p>\n<p>I personally prefer to blame the trouble on the notion of space, rather than the axiom of choice. As far as possible, geometric problems should be the business of geometry, not logic or set theory. Mathematicians are used to operating with various kinds of spaces (in geometry, in analysis, in topology, in algebraic geometry, in computation, etc.) and so it seems only natural that one should worry about using the correct notion of space first, and about underlying foundational principles later. Good math is immune to changes in foundations.</p>\n" }, { "answer_id": 260065, "score": 42, "text": "<p>It's notable that most of the \"bread and butter\" mathematical consequences of the axiom of choice are actually consequences of <em>countable</em> choice. (Every infinite set contains a countable subset, a countable union of countable sets is countable, etc.) The Hahn-Banach theorem is a counterexample, but only if you want it for nonseparable spaces, and I can't think of any time I've ever needed this. When restricted to separable Banach spaces it doesn't require any choice principle at all! Whereas the seemingly pathological consequences of choice (existence of nonmeasurable sets, Banach-Tarski, well-ordering of the real line) generally do not follow from countable choice.</p>\n\n<p>So the argument from mathematical value seems to me to support countable choice more than full choice. But that isn't a very strong argument, is it? We can't decide whether an axiom is true based on whether we like its consequences. At best it's suggestive.</p>\n\n<p>Incidentally, I had the impression when I read Zermelo that he had great polemical skill, but none of his arguments seemed to get directly to the <em>truth</em> of the axiom. He argues for the mathematical value of the axiom. He points out that his critics have themselves on occasion unwittingly used the axiom, which is a devastating point, but has little bearing on the question of truth. (If I'm not mistaken, those unwitting uses were all of countable choice, by the way.)</p>\n\n<p>You ask if one should \"believe\" the axiom of choice, and I think you are right to put the word \"believe\" in quotes. I feel strongly that set-theoretic assertions are objectively meaningful, but I also feel that philosophers of mathematics have done a very poor job of clarifying what sets are. (Halmos: \"A pack of wolves, a bunch of grapes, or a flock of pigeons are all examples of sets of things.\" Black: \"It ought then to make sense, at least sometimes, to speak of being pursued by a set, or eating a set, or putting a set to flight.\") If we can't even get that straight, it's hard to come to grips with questions about the truth of questionable axioms.</p>\n" }, { "answer_id": 260120, "score": 36, "text": "<p>The other answers don't seem to have said much about why the axiom of choice is widely regarded as plausible. Let me try to address that question.</p>\n\n<p>First let's dispose of some non-reasons. In response to your questions, I don't know of anyone who thinks that the Banach&ndash;Tarski paradox is a reason to believe in the axiom of choice. I also don't know of anyone who argues, \"It is a priori plausible that there exist non-measurable sets; so the fact that the axiom of choice yields the attractive conclusion that there are non-measurable sets is a point in favor of believing the axiom of choice.\" Instead, those who are comfortable with the existence of non-measurable sets typically start by accepting the axiom of choice, and then they accept non-measurable sets as \"part of the territory\" that comes with the axiom of choice.</p>\n\n<p>Those who think that Banach&ndash;Tarski casts doubt on the axiom of choice typically have a philosophical predisposition that <i>math is supposed to model the physical world closely</i>. So for example, $\\mathbb R^3$ is not just a random mathematical structure that we study purely for its own sake; it is supposed to be a decent model of physical space (or at least, open subsets of $\\mathbb R^3$ are supposed to model localized regions of physical space). Banach&ndash;Tarski, when given a direct physical interpretation in this way, yields something that we \"know\" makes no physical sense, and so if we think that math is supposed to yield physical truth in this way, then Banach&ndash;Tarski is going to lead us to reject something in the math. Whether that \"something\" we reject is the axiom of choice is a separate question, and Andrej Bauer's excellent answer shows that there are other options, but the point I want to highlight is that we're going to be led down this path in the first place only if we have certain presuppositions about how math and physics are supposed to relate.</p>\n\n<p>There are others who don't view set theory in this way. According to them, set theory is supposed to be about <i>abstract collections of things</i>, and the way to arrive at axioms is by abstractly thinking about what properties they should have, not by comparing them with the physical world. The axiom of choice can be thought of as saying that if you have a bunch of nonempty collections of things, then there is another collection of things that contains one element from each of your original nonempty collections. Stated this way, the principle sounds intuitively plausible, and I would argue that this intuitive plausibility is, at least implicitly, the main argument in the minds of most people who accept the axiom of choice. If this is the way you think, then non-measurable sets and Banach&ndash;Tarski are not going to dissuade you from accepting the axiom of choice. Those phenomena will just lead you to say that we can't arrive at physical predictions from mathematics in such a naive manner; instead, to do physics, we have to formulate <i>physical theories</i>. Math can of course help a lot with the construction of physical theories, but it's not as simple as just saying that the mathematical theory of $\\mathbb R^3$ <i>is</i> our theory of physical space.</p>\n\n<p>These two options aren't the only options. The <a href=\"https://en.wikipedia.org/wiki/Solovay_model\">work of Solovay</a> shows that you can, to a large extent, <a href=\"https://books.google.com/ngrams/graph?content=have+your+cake+and+eat%2Ceat+your+cake+and+have&amp;year_start=1840&amp;year_end=2000&amp;corpus=15&amp;smoothing=3&amp;share=&amp;direct_url=t1%3B%2Chave%20your%20cake%20and%20eat%3B%2Cc0%3B.t1%3B%2Ceat%20your%20cake%20and%20have%3B%2Cc0\">have your cake and eat it too</a>, by working in a set-theoretic universe where all subsets of $\\mathbb R^n$ are Lebesgue-measurable and a weakened, but still quite strong, version of the axiom of choice known as \"dependent choice\" is available. Why Solovay's model hasn't become more popular is not completely clear, but perhaps part of the reason is that it feels like a \"compromise position,\" and the people in the two different camps above have not seen any need to migrate to that kind of compromise.</p>\n" }, { "answer_id": 260118, "score": 27, "text": "<blockquote>\n <p>Are there reasons why it is plausible (for physicists, philosophers,\n mathematicians) to believe that not all sets should be measurable?</p>\n</blockquote>\n\n<p>Yes. If every set of reals is Lebesgue measurable, then you can partition $\\mathbb{R}$ into <em>more than continuum many</em> pairwise disjoint non-empty pieces. (See <a href=\"https://mathoverflow.net/a/22935/33039\">this answer and comments</a> for the details.)</p>\n\n<p>Surely the Banach-Tarski paradox seems unintuitive. But having a set that can be broken up into more pieces than there originally were... is just wrong.</p>\n" }, { "answer_id": 260092, "score": 18, "text": "<p>Arguments from physics may not help. Here is Bryce DeWitt <a href=\"http://science.sciencemag.org/content/182/4113/705\">reviewing</a> Stephen Hawking and G.F.R. Ellis using the axiom of choice in 1973:</p>\n\n<blockquote>\n <p>The book also contains one failure to distinguish between mathematics\n and physics that is actually serious. This is in the proof of the\n main theorem of chapter 7, that given a set of Cauchy data on a smooth\n spacelike hypersurface there exists a unique maximal development\n therefrom of Einstein’s empty-space equations. The proof, essentially\n due to Choquet-Bruhat and Geroch, makes use of the axiom of choice,\n in the guise of Zorn’s lemma. Now mathematicians may use this axiom if\n they wish, but it has no place in physics. Physicists are already\n stretching things, from an operational standpoint, in using the axiom\n of infinity.</p>\n \n <p>It is not a question here of resurrecting an old and out-of-date\n mathematical controversy. The simple fact is that the axiom of\n choice never is really needed except when dealing with sets and\n relations in non-constructive ways. Many remarkable and beautiful\n theorems can be proved only with its aid. But its irrelevance to\n physics should be evident from the fact that its denial, as Paul\n Cohen has shown us, is equally consistent with the other axioms\n of set theory. And these other axioms suffice for the constructions of\n the real numbers, Hilbert spaces, C* algebras, and pseudo-Riemannian\n manifolds–that is, of all the paraphernalia of theoretical physics.</p>\n \n <p>In “proving” the global Cauchy development theorem with the aid of\n Zorn’s lemma what one is actually doing is assuming that a “choice\n function” exists for every set of developments extending a given\n Cauchy development. This, of course, is begging the question. The\n physicist’s job is not done until he can show, by an explicit\n algorithm or construction, how one could in principle always select a\n member from every such set of developments. Failing that he has proved\n nothing.</p>\n</blockquote>\n\n<p>Some physicists want to use the axiom of choice, but some physicists don't.</p>\n" }, { "answer_id": 260073, "score": 13, "text": "<p>Just a quick addendum: the result was actually found by Felix Hausdorff and then repackaged in a more spectacular form by Banach and Tarski. Hausdorff' point was precisely to show that the axiom of choice leads to such unreasonable consequences that it should probably be avoided. This was a hot topic of discussion during the following decades, and it seems that the answer at large from the mathematical community is: ok, there are inconveniences, but the advantages of using the AC are superior and we prefer to have it available.</p>\n" }, { "answer_id": 260128, "score": 9, "text": "<p>I think one of the strongest arguments for the axiom of choice is that every model of ZF contains as an inner model a constructible universe $L$, and $AC$ is a theorem of the constructible universe. We have</p>\n\n<p>$$ ZF+(V=L) \\vdash AC $$</p>\n\n<p>In other words, a necessary condition for asserting $\\neg AC$ is to first assert \"there exists a set that cannot be constructed\" &mdash; that is, it requires positing the existence of additional structure above and beyond what is guaranteed by ZF.</p>\n\n<p>Thus, it seems clear to me that ZFC is by far the better choice for <em>foundations</em>. One may still wish to work with another set theoretic universe, but that's most appropriately done as the study of <em>additional structure</em> built atop of the foundations, not by rewriting the foundations themselves.</p>\n" }, { "answer_id": 260141, "score": 8, "text": "<p>Physical applications of the Banach–Tarski theorem were explored by <a href=\"https://archive.org/details/Startling_Stories_v18n03_1949-01\" rel=\"noreferrer\">Henry Kuttner, The Time Axis, <em>Startling Stories</em> 18:3 (January 1949), 13–82</a>. Some excerpts from <a href=\"https://archive.org/stream/Startling_Stories_v18n03_1949-01#page/n65/mode/2up\" rel=\"noreferrer\">pp. 66–67</a>:</p>\n\n<blockquote>\n <p>\"'Professor Raphael M. Robinson of the University of California now shows that it is possible to divide a solid sphere into a minimum of five pieces and reassemble them to form two spheres of the same size as the original one. Two of the pieces are used to form one of the new spheres and three to form the other.<br>\n <br>\"'Some of the pieces must necessarily be of such complicated structure that it is impossible to assign volume to them. Otherwise the sum of the volumes of the five pieces would have to be equal both to the volume of the original sphere and to the sum of the volumes of the two new spheres, which is twice as great.'\"<br>\n <br>[. . . .]<br>\n <br>\"This is it,\" he said.<br>\n <br>Even the crowd around the neural-web table thinned as the workers in the laboratory flocked around him to watch.<br>\n <br>He had a sphere about the size of a grapefruit, floating in mid-air above his table. He did things to it with quick flashes of light that acted exactly like knives, in that it fell apart wherever the lights touched, but I got the impression that those divisions were much less simple than knife-cuts would be. The light shivered as it slashed and the cuts must have been very complex, dividing molecules with a selective precision beyond my powers of comprehension.<br>\n <br>The sphere floated apart. It changed shape under the lights. I am pretty sure it changed shape in four dimensions, because after a while I literally could not watch any more. The shape did agonizing things to my eyes when I tried to focus on it.<br>\n <br>When I heard a long sigh go up simultaneously from the watchers I risked a look again.<br>\n <br>There were two spheres floating where one had floated before.</p>\n</blockquote>\n" }, { "answer_id": 260680, "score": 5, "text": "<p>I think it is interesting to re-phrase this question relative to <strong>other</strong> axioms and/or theorems of ZFC. What Andrej Bauer's answer suggests is that it may not be the axiom of choice <em>per se</em> that is the culprit, but rather the underlying structure.</p>\n\n<p>For example, it is provable that the existence of non-Lebesgue measurable sets and the Banach–Tarski paradox are <strong>both</strong> an implication over ZF (without Choice) of the Hahn–Banach theorem (HB) in functional analysis. This means that we can prove those \"pathologies\" as a theorems from ZF+HB. </p>\n\n<p>Another way to put it is to analyze the structure within the context of different orders of logic. In ordinary first-order logic Choice is provably equivalent to the well-ordering theorem (WO) over ZF. What is different in second-order logic is that WO is strictly stronger than choice: WO $\\vdash_{ZF}$ Choice, but Choice $\\nvdash_{ZF}$ WO.</p>\n\n<p>In other words, one may get the feeling that perhaps there has been too much emphasis put on Choice. Other axioms and theorems can prove just as problematic. The approach described above is the essence of Reverse Mathematics (RM). While RM has been traditionally carried out at a much lower proof-theoretic strength level (subsystems of second-order arithmetic), in my opinion this provides a very useful framework for analyzing the foundations of other parts of mathematics.</p>\n" }, { "answer_id": 260072, "score": 120, "text": "<p>There are <em>two</em> ingredients in the Banach-Tarski decomposition theorem:</p>\n<ol>\n<li>The notion of <em>space</em>, together with derived notions of <em>part</em> and <em>decomposition</em>.</li>\n<li>The axiom of choice.</li>\n</ol>\n<p>Most discussion about the theorem revolve around the axiom of choice. I would like to point out that the notion of space can be put under scrutiny as well.</p>\n<p>The Banach-Tarski decomposition of the sphere produces non-measurable parts of the sphere. If we <em>restrict</em> the notion of &quot;part&quot; to &quot;measurable subset&quot; the theorem disappears. For instance, if we move over into a model of set theory (without choice) in which all sets are measurable, we will have no Banach-Tarski. This is all well known.</p>\n<p>Somewhat amazingly, we can make the Banach-Tarski decomposition go away by <em>extending</em> the notion of subspace, and keep choice too. Alex Simpson in <a href=\"https://web.archive.org/web/20160909042100/http://homepages.inf.ed.ac.uk/als/Research/Sources/mrs.pdf\" rel=\"noreferrer\">Measure, Randomness and Sublocales</a> (Annals of Pure and Applied Logic, 163(11), pp. 1642-1659, 2012) shows that this is achieved by generalizing the notion of topological space to that of <a href=\"https://ncatlab.org/nlab/show/locale\" rel=\"noreferrer\">locale</a>. He explains it thus:</p>\n<blockquote>\n<p>&quot;The different pieces in the partitions defined by Vitali and by Banach and Tarski are deeply intertangled with each other. According to our notion of “part”, two such intertangled pieces are not disjoint from each other, so additivity does not apply. An intuitive explanation for the failure of disjointness is that, although two such pieces share no point in common, they nevertheless overlap on the topological “glue” that bonds neighbouring points in <span class=\"math-container\">$\\mathbb{R}^n$</span> together.&quot;</p>\n</blockquote>\n<p>Peter Johnstone explained in <a href=\"http://pointlesstopology.com/the-point-of-pointless-topology.pdf\" rel=\"noreferrer\">The point of pointless topology</a> why locales have mathematical significance that goes far beyond fixing a strange theorem about decomposition of the sphere. Why isn't everyone using locales? I do not know, I think it is purely a historic accident. At some point in the 20th century mathematicians collectively lost the idea that there is more to space than just its points.</p>\n<p>I personally prefer to blame the trouble on the notion of space, rather than the axiom of choice. As far as possible, geometric problems should be the business of geometry, not logic or set theory. Mathematicians are used to operating with various kinds of spaces (in geometry, in analysis, in topology, in algebraic geometry, in computation, etc.) and so it seems only natural that one should worry about using the correct notion of space first, and about underlying foundational principles later. Good math is immune to changes in foundations.</p>\n" }, { "answer_id": 260065, "score": 42, "text": "<p>It's notable that most of the \"bread and butter\" mathematical consequences of the axiom of choice are actually consequences of <em>countable</em> choice. (Every infinite set contains a countable subset, a countable union of countable sets is countable, etc.) The Hahn-Banach theorem is a counterexample, but only if you want it for nonseparable spaces, and I can't think of any time I've ever needed this. When restricted to separable Banach spaces it doesn't require any choice principle at all! Whereas the seemingly pathological consequences of choice (existence of nonmeasurable sets, Banach-Tarski, well-ordering of the real line) generally do not follow from countable choice.</p>\n\n<p>So the argument from mathematical value seems to me to support countable choice more than full choice. But that isn't a very strong argument, is it? We can't decide whether an axiom is true based on whether we like its consequences. At best it's suggestive.</p>\n\n<p>Incidentally, I had the impression when I read Zermelo that he had great polemical skill, but none of his arguments seemed to get directly to the <em>truth</em> of the axiom. He argues for the mathematical value of the axiom. He points out that his critics have themselves on occasion unwittingly used the axiom, which is a devastating point, but has little bearing on the question of truth. (If I'm not mistaken, those unwitting uses were all of countable choice, by the way.)</p>\n\n<p>You ask if one should \"believe\" the axiom of choice, and I think you are right to put the word \"believe\" in quotes. I feel strongly that set-theoretic assertions are objectively meaningful, but I also feel that philosophers of mathematics have done a very poor job of clarifying what sets are. (Halmos: \"A pack of wolves, a bunch of grapes, or a flock of pigeons are all examples of sets of things.\" Black: \"It ought then to make sense, at least sometimes, to speak of being pursued by a set, or eating a set, or putting a set to flight.\") If we can't even get that straight, it's hard to come to grips with questions about the truth of questionable axioms.</p>\n" }, { "answer_id": 260120, "score": 36, "text": "<p>The other answers don't seem to have said much about why the axiom of choice is widely regarded as plausible. Let me try to address that question.</p>\n\n<p>First let's dispose of some non-reasons. In response to your questions, I don't know of anyone who thinks that the Banach&ndash;Tarski paradox is a reason to believe in the axiom of choice. I also don't know of anyone who argues, \"It is a priori plausible that there exist non-measurable sets; so the fact that the axiom of choice yields the attractive conclusion that there are non-measurable sets is a point in favor of believing the axiom of choice.\" Instead, those who are comfortable with the existence of non-measurable sets typically start by accepting the axiom of choice, and then they accept non-measurable sets as \"part of the territory\" that comes with the axiom of choice.</p>\n\n<p>Those who think that Banach&ndash;Tarski casts doubt on the axiom of choice typically have a philosophical predisposition that <i>math is supposed to model the physical world closely</i>. So for example, $\\mathbb R^3$ is not just a random mathematical structure that we study purely for its own sake; it is supposed to be a decent model of physical space (or at least, open subsets of $\\mathbb R^3$ are supposed to model localized regions of physical space). Banach&ndash;Tarski, when given a direct physical interpretation in this way, yields something that we \"know\" makes no physical sense, and so if we think that math is supposed to yield physical truth in this way, then Banach&ndash;Tarski is going to lead us to reject something in the math. Whether that \"something\" we reject is the axiom of choice is a separate question, and Andrej Bauer's excellent answer shows that there are other options, but the point I want to highlight is that we're going to be led down this path in the first place only if we have certain presuppositions about how math and physics are supposed to relate.</p>\n\n<p>There are others who don't view set theory in this way. According to them, set theory is supposed to be about <i>abstract collections of things</i>, and the way to arrive at axioms is by abstractly thinking about what properties they should have, not by comparing them with the physical world. The axiom of choice can be thought of as saying that if you have a bunch of nonempty collections of things, then there is another collection of things that contains one element from each of your original nonempty collections. Stated this way, the principle sounds intuitively plausible, and I would argue that this intuitive plausibility is, at least implicitly, the main argument in the minds of most people who accept the axiom of choice. If this is the way you think, then non-measurable sets and Banach&ndash;Tarski are not going to dissuade you from accepting the axiom of choice. Those phenomena will just lead you to say that we can't arrive at physical predictions from mathematics in such a naive manner; instead, to do physics, we have to formulate <i>physical theories</i>. Math can of course help a lot with the construction of physical theories, but it's not as simple as just saying that the mathematical theory of $\\mathbb R^3$ <i>is</i> our theory of physical space.</p>\n\n<p>These two options aren't the only options. The <a href=\"https://en.wikipedia.org/wiki/Solovay_model\">work of Solovay</a> shows that you can, to a large extent, <a href=\"https://books.google.com/ngrams/graph?content=have+your+cake+and+eat%2Ceat+your+cake+and+have&amp;year_start=1840&amp;year_end=2000&amp;corpus=15&amp;smoothing=3&amp;share=&amp;direct_url=t1%3B%2Chave%20your%20cake%20and%20eat%3B%2Cc0%3B.t1%3B%2Ceat%20your%20cake%20and%20have%3B%2Cc0\">have your cake and eat it too</a>, by working in a set-theoretic universe where all subsets of $\\mathbb R^n$ are Lebesgue-measurable and a weakened, but still quite strong, version of the axiom of choice known as \"dependent choice\" is available. Why Solovay's model hasn't become more popular is not completely clear, but perhaps part of the reason is that it feels like a \"compromise position,\" and the people in the two different camps above have not seen any need to migrate to that kind of compromise.</p>\n" }, { "answer_id": 260118, "score": 27, "text": "<blockquote>\n <p>Are there reasons why it is plausible (for physicists, philosophers,\n mathematicians) to believe that not all sets should be measurable?</p>\n</blockquote>\n\n<p>Yes. If every set of reals is Lebesgue measurable, then you can partition $\\mathbb{R}$ into <em>more than continuum many</em> pairwise disjoint non-empty pieces. (See <a href=\"https://mathoverflow.net/a/22935/33039\">this answer and comments</a> for the details.)</p>\n\n<p>Surely the Banach-Tarski paradox seems unintuitive. But having a set that can be broken up into more pieces than there originally were... is just wrong.</p>\n" }, { "answer_id": 260092, "score": 18, "text": "<p>Arguments from physics may not help. Here is Bryce DeWitt <a href=\"http://science.sciencemag.org/content/182/4113/705\">reviewing</a> Stephen Hawking and G.F.R. Ellis using the axiom of choice in 1973:</p>\n\n<blockquote>\n <p>The book also contains one failure to distinguish between mathematics\n and physics that is actually serious. This is in the proof of the\n main theorem of chapter 7, that given a set of Cauchy data on a smooth\n spacelike hypersurface there exists a unique maximal development\n therefrom of Einstein’s empty-space equations. The proof, essentially\n due to Choquet-Bruhat and Geroch, makes use of the axiom of choice,\n in the guise of Zorn’s lemma. Now mathematicians may use this axiom if\n they wish, but it has no place in physics. Physicists are already\n stretching things, from an operational standpoint, in using the axiom\n of infinity.</p>\n \n <p>It is not a question here of resurrecting an old and out-of-date\n mathematical controversy. The simple fact is that the axiom of\n choice never is really needed except when dealing with sets and\n relations in non-constructive ways. Many remarkable and beautiful\n theorems can be proved only with its aid. But its irrelevance to\n physics should be evident from the fact that its denial, as Paul\n Cohen has shown us, is equally consistent with the other axioms\n of set theory. And these other axioms suffice for the constructions of\n the real numbers, Hilbert spaces, C* algebras, and pseudo-Riemannian\n manifolds–that is, of all the paraphernalia of theoretical physics.</p>\n \n <p>In “proving” the global Cauchy development theorem with the aid of\n Zorn’s lemma what one is actually doing is assuming that a “choice\n function” exists for every set of developments extending a given\n Cauchy development. This, of course, is begging the question. The\n physicist’s job is not done until he can show, by an explicit\n algorithm or construction, how one could in principle always select a\n member from every such set of developments. Failing that he has proved\n nothing.</p>\n</blockquote>\n\n<p>Some physicists want to use the axiom of choice, but some physicists don't.</p>\n" }, { "answer_id": 260073, "score": 13, "text": "<p>Just a quick addendum: the result was actually found by Felix Hausdorff and then repackaged in a more spectacular form by Banach and Tarski. Hausdorff' point was precisely to show that the axiom of choice leads to such unreasonable consequences that it should probably be avoided. This was a hot topic of discussion during the following decades, and it seems that the answer at large from the mathematical community is: ok, there are inconveniences, but the advantages of using the AC are superior and we prefer to have it available.</p>\n" }, { "answer_id": 260128, "score": 9, "text": "<p>I think one of the strongest arguments for the axiom of choice is that every model of ZF contains as an inner model a constructible universe $L$, and $AC$ is a theorem of the constructible universe. We have</p>\n\n<p>$$ ZF+(V=L) \\vdash AC $$</p>\n\n<p>In other words, a necessary condition for asserting $\\neg AC$ is to first assert \"there exists a set that cannot be constructed\" &mdash; that is, it requires positing the existence of additional structure above and beyond what is guaranteed by ZF.</p>\n\n<p>Thus, it seems clear to me that ZFC is by far the better choice for <em>foundations</em>. One may still wish to work with another set theoretic universe, but that's most appropriately done as the study of <em>additional structure</em> built atop of the foundations, not by rewriting the foundations themselves.</p>\n" }, { "answer_id": 260141, "score": 8, "text": "<p>Physical applications of the Banach–Tarski theorem were explored by <a href=\"https://archive.org/details/Startling_Stories_v18n03_1949-01\" rel=\"noreferrer\">Henry Kuttner, The Time Axis, <em>Startling Stories</em> 18:3 (January 1949), 13–82</a>. Some excerpts from <a href=\"https://archive.org/stream/Startling_Stories_v18n03_1949-01#page/n65/mode/2up\" rel=\"noreferrer\">pp. 66–67</a>:</p>\n\n<blockquote>\n <p>\"'Professor Raphael M. Robinson of the University of California now shows that it is possible to divide a solid sphere into a minimum of five pieces and reassemble them to form two spheres of the same size as the original one. Two of the pieces are used to form one of the new spheres and three to form the other.<br>\n <br>\"'Some of the pieces must necessarily be of such complicated structure that it is impossible to assign volume to them. Otherwise the sum of the volumes of the five pieces would have to be equal both to the volume of the original sphere and to the sum of the volumes of the two new spheres, which is twice as great.'\"<br>\n <br>[. . . .]<br>\n <br>\"This is it,\" he said.<br>\n <br>Even the crowd around the neural-web table thinned as the workers in the laboratory flocked around him to watch.<br>\n <br>He had a sphere about the size of a grapefruit, floating in mid-air above his table. He did things to it with quick flashes of light that acted exactly like knives, in that it fell apart wherever the lights touched, but I got the impression that those divisions were much less simple than knife-cuts would be. The light shivered as it slashed and the cuts must have been very complex, dividing molecules with a selective precision beyond my powers of comprehension.<br>\n <br>The sphere floated apart. It changed shape under the lights. I am pretty sure it changed shape in four dimensions, because after a while I literally could not watch any more. The shape did agonizing things to my eyes when I tried to focus on it.<br>\n <br>When I heard a long sigh go up simultaneously from the watchers I risked a look again.<br>\n <br>There were two spheres floating where one had floated before.</p>\n</blockquote>\n" }, { "answer_id": 260680, "score": 5, "text": "<p>I think it is interesting to re-phrase this question relative to <strong>other</strong> axioms and/or theorems of ZFC. What Andrej Bauer's answer suggests is that it may not be the axiom of choice <em>per se</em> that is the culprit, but rather the underlying structure.</p>\n\n<p>For example, it is provable that the existence of non-Lebesgue measurable sets and the Banach–Tarski paradox are <strong>both</strong> an implication over ZF (without Choice) of the Hahn–Banach theorem (HB) in functional analysis. This means that we can prove those \"pathologies\" as a theorems from ZF+HB. </p>\n\n<p>Another way to put it is to analyze the structure within the context of different orders of logic. In ordinary first-order logic Choice is provably equivalent to the well-ordering theorem (WO) over ZF. What is different in second-order logic is that WO is strictly stronger than choice: WO $\\vdash_{ZF}$ Choice, but Choice $\\nvdash_{ZF}$ WO.</p>\n\n<p>In other words, one may get the feeling that perhaps there has been too much emphasis put on Choice. Other axioms and theorems can prove just as problematic. The approach described above is the essence of Reverse Mathematics (RM). While RM has been traditionally carried out at a much lower proof-theoretic strength level (subsystems of second-order arithmetic), in my opinion this provides a very useful framework for analyzing the foundations of other parts of mathematics.</p>\n" }, { "answer_id": 260072, "score": 120, "text": "<p>There are <em>two</em> ingredients in the Banach-Tarski decomposition theorem:</p>\n<ol>\n<li>The notion of <em>space</em>, together with derived notions of <em>part</em> and <em>decomposition</em>.</li>\n<li>The axiom of choice.</li>\n</ol>\n<p>Most discussion about the theorem revolve around the axiom of choice. I would like to point out that the notion of space can be put under scrutiny as well.</p>\n<p>The Banach-Tarski decomposition of the sphere produces non-measurable parts of the sphere. If we <em>restrict</em> the notion of &quot;part&quot; to &quot;measurable subset&quot; the theorem disappears. For instance, if we move over into a model of set theory (without choice) in which all sets are measurable, we will have no Banach-Tarski. This is all well known.</p>\n<p>Somewhat amazingly, we can make the Banach-Tarski decomposition go away by <em>extending</em> the notion of subspace, and keep choice too. Alex Simpson in <a href=\"https://web.archive.org/web/20160909042100/http://homepages.inf.ed.ac.uk/als/Research/Sources/mrs.pdf\" rel=\"noreferrer\">Measure, Randomness and Sublocales</a> (Annals of Pure and Applied Logic, 163(11), pp. 1642-1659, 2012) shows that this is achieved by generalizing the notion of topological space to that of <a href=\"https://ncatlab.org/nlab/show/locale\" rel=\"noreferrer\">locale</a>. He explains it thus:</p>\n<blockquote>\n<p>&quot;The different pieces in the partitions defined by Vitali and by Banach and Tarski are deeply intertangled with each other. According to our notion of “part”, two such intertangled pieces are not disjoint from each other, so additivity does not apply. An intuitive explanation for the failure of disjointness is that, although two such pieces share no point in common, they nevertheless overlap on the topological “glue” that bonds neighbouring points in <span class=\"math-container\">$\\mathbb{R}^n$</span> together.&quot;</p>\n</blockquote>\n<p>Peter Johnstone explained in <a href=\"http://pointlesstopology.com/the-point-of-pointless-topology.pdf\" rel=\"noreferrer\">The point of pointless topology</a> why locales have mathematical significance that goes far beyond fixing a strange theorem about decomposition of the sphere. Why isn't everyone using locales? I do not know, I think it is purely a historic accident. At some point in the 20th century mathematicians collectively lost the idea that there is more to space than just its points.</p>\n<p>I personally prefer to blame the trouble on the notion of space, rather than the axiom of choice. As far as possible, geometric problems should be the business of geometry, not logic or set theory. Mathematicians are used to operating with various kinds of spaces (in geometry, in analysis, in topology, in algebraic geometry, in computation, etc.) and so it seems only natural that one should worry about using the correct notion of space first, and about underlying foundational principles later. Good math is immune to changes in foundations.</p>\n" }, { "answer_id": 260065, "score": 42, "text": "<p>It's notable that most of the \"bread and butter\" mathematical consequences of the axiom of choice are actually consequences of <em>countable</em> choice. (Every infinite set contains a countable subset, a countable union of countable sets is countable, etc.) The Hahn-Banach theorem is a counterexample, but only if you want it for nonseparable spaces, and I can't think of any time I've ever needed this. When restricted to separable Banach spaces it doesn't require any choice principle at all! Whereas the seemingly pathological consequences of choice (existence of nonmeasurable sets, Banach-Tarski, well-ordering of the real line) generally do not follow from countable choice.</p>\n\n<p>So the argument from mathematical value seems to me to support countable choice more than full choice. But that isn't a very strong argument, is it? We can't decide whether an axiom is true based on whether we like its consequences. At best it's suggestive.</p>\n\n<p>Incidentally, I had the impression when I read Zermelo that he had great polemical skill, but none of his arguments seemed to get directly to the <em>truth</em> of the axiom. He argues for the mathematical value of the axiom. He points out that his critics have themselves on occasion unwittingly used the axiom, which is a devastating point, but has little bearing on the question of truth. (If I'm not mistaken, those unwitting uses were all of countable choice, by the way.)</p>\n\n<p>You ask if one should \"believe\" the axiom of choice, and I think you are right to put the word \"believe\" in quotes. I feel strongly that set-theoretic assertions are objectively meaningful, but I also feel that philosophers of mathematics have done a very poor job of clarifying what sets are. (Halmos: \"A pack of wolves, a bunch of grapes, or a flock of pigeons are all examples of sets of things.\" Black: \"It ought then to make sense, at least sometimes, to speak of being pursued by a set, or eating a set, or putting a set to flight.\") If we can't even get that straight, it's hard to come to grips with questions about the truth of questionable axioms.</p>\n" }, { "answer_id": 260120, "score": 36, "text": "<p>The other answers don't seem to have said much about why the axiom of choice is widely regarded as plausible. Let me try to address that question.</p>\n\n<p>First let's dispose of some non-reasons. In response to your questions, I don't know of anyone who thinks that the Banach&ndash;Tarski paradox is a reason to believe in the axiom of choice. I also don't know of anyone who argues, \"It is a priori plausible that there exist non-measurable sets; so the fact that the axiom of choice yields the attractive conclusion that there are non-measurable sets is a point in favor of believing the axiom of choice.\" Instead, those who are comfortable with the existence of non-measurable sets typically start by accepting the axiom of choice, and then they accept non-measurable sets as \"part of the territory\" that comes with the axiom of choice.</p>\n\n<p>Those who think that Banach&ndash;Tarski casts doubt on the axiom of choice typically have a philosophical predisposition that <i>math is supposed to model the physical world closely</i>. So for example, $\\mathbb R^3$ is not just a random mathematical structure that we study purely for its own sake; it is supposed to be a decent model of physical space (or at least, open subsets of $\\mathbb R^3$ are supposed to model localized regions of physical space). Banach&ndash;Tarski, when given a direct physical interpretation in this way, yields something that we \"know\" makes no physical sense, and so if we think that math is supposed to yield physical truth in this way, then Banach&ndash;Tarski is going to lead us to reject something in the math. Whether that \"something\" we reject is the axiom of choice is a separate question, and Andrej Bauer's excellent answer shows that there are other options, but the point I want to highlight is that we're going to be led down this path in the first place only if we have certain presuppositions about how math and physics are supposed to relate.</p>\n\n<p>There are others who don't view set theory in this way. According to them, set theory is supposed to be about <i>abstract collections of things</i>, and the way to arrive at axioms is by abstractly thinking about what properties they should have, not by comparing them with the physical world. The axiom of choice can be thought of as saying that if you have a bunch of nonempty collections of things, then there is another collection of things that contains one element from each of your original nonempty collections. Stated this way, the principle sounds intuitively plausible, and I would argue that this intuitive plausibility is, at least implicitly, the main argument in the minds of most people who accept the axiom of choice. If this is the way you think, then non-measurable sets and Banach&ndash;Tarski are not going to dissuade you from accepting the axiom of choice. Those phenomena will just lead you to say that we can't arrive at physical predictions from mathematics in such a naive manner; instead, to do physics, we have to formulate <i>physical theories</i>. Math can of course help a lot with the construction of physical theories, but it's not as simple as just saying that the mathematical theory of $\\mathbb R^3$ <i>is</i> our theory of physical space.</p>\n\n<p>These two options aren't the only options. The <a href=\"https://en.wikipedia.org/wiki/Solovay_model\">work of Solovay</a> shows that you can, to a large extent, <a href=\"https://books.google.com/ngrams/graph?content=have+your+cake+and+eat%2Ceat+your+cake+and+have&amp;year_start=1840&amp;year_end=2000&amp;corpus=15&amp;smoothing=3&amp;share=&amp;direct_url=t1%3B%2Chave%20your%20cake%20and%20eat%3B%2Cc0%3B.t1%3B%2Ceat%20your%20cake%20and%20have%3B%2Cc0\">have your cake and eat it too</a>, by working in a set-theoretic universe where all subsets of $\\mathbb R^n$ are Lebesgue-measurable and a weakened, but still quite strong, version of the axiom of choice known as \"dependent choice\" is available. Why Solovay's model hasn't become more popular is not completely clear, but perhaps part of the reason is that it feels like a \"compromise position,\" and the people in the two different camps above have not seen any need to migrate to that kind of compromise.</p>\n" }, { "answer_id": 260118, "score": 27, "text": "<blockquote>\n <p>Are there reasons why it is plausible (for physicists, philosophers,\n mathematicians) to believe that not all sets should be measurable?</p>\n</blockquote>\n\n<p>Yes. If every set of reals is Lebesgue measurable, then you can partition $\\mathbb{R}$ into <em>more than continuum many</em> pairwise disjoint non-empty pieces. (See <a href=\"https://mathoverflow.net/a/22935/33039\">this answer and comments</a> for the details.)</p>\n\n<p>Surely the Banach-Tarski paradox seems unintuitive. But having a set that can be broken up into more pieces than there originally were... is just wrong.</p>\n" }, { "answer_id": 260092, "score": 18, "text": "<p>Arguments from physics may not help. Here is Bryce DeWitt <a href=\"http://science.sciencemag.org/content/182/4113/705\">reviewing</a> Stephen Hawking and G.F.R. Ellis using the axiom of choice in 1973:</p>\n\n<blockquote>\n <p>The book also contains one failure to distinguish between mathematics\n and physics that is actually serious. This is in the proof of the\n main theorem of chapter 7, that given a set of Cauchy data on a smooth\n spacelike hypersurface there exists a unique maximal development\n therefrom of Einstein’s empty-space equations. The proof, essentially\n due to Choquet-Bruhat and Geroch, makes use of the axiom of choice,\n in the guise of Zorn’s lemma. Now mathematicians may use this axiom if\n they wish, but it has no place in physics. Physicists are already\n stretching things, from an operational standpoint, in using the axiom\n of infinity.</p>\n \n <p>It is not a question here of resurrecting an old and out-of-date\n mathematical controversy. The simple fact is that the axiom of\n choice never is really needed except when dealing with sets and\n relations in non-constructive ways. Many remarkable and beautiful\n theorems can be proved only with its aid. But its irrelevance to\n physics should be evident from the fact that its denial, as Paul\n Cohen has shown us, is equally consistent with the other axioms\n of set theory. And these other axioms suffice for the constructions of\n the real numbers, Hilbert spaces, C* algebras, and pseudo-Riemannian\n manifolds–that is, of all the paraphernalia of theoretical physics.</p>\n \n <p>In “proving” the global Cauchy development theorem with the aid of\n Zorn’s lemma what one is actually doing is assuming that a “choice\n function” exists for every set of developments extending a given\n Cauchy development. This, of course, is begging the question. The\n physicist’s job is not done until he can show, by an explicit\n algorithm or construction, how one could in principle always select a\n member from every such set of developments. Failing that he has proved\n nothing.</p>\n</blockquote>\n\n<p>Some physicists want to use the axiom of choice, but some physicists don't.</p>\n" }, { "answer_id": 260073, "score": 13, "text": "<p>Just a quick addendum: the result was actually found by Felix Hausdorff and then repackaged in a more spectacular form by Banach and Tarski. Hausdorff' point was precisely to show that the axiom of choice leads to such unreasonable consequences that it should probably be avoided. This was a hot topic of discussion during the following decades, and it seems that the answer at large from the mathematical community is: ok, there are inconveniences, but the advantages of using the AC are superior and we prefer to have it available.</p>\n" }, { "answer_id": 260128, "score": 9, "text": "<p>I think one of the strongest arguments for the axiom of choice is that every model of ZF contains as an inner model a constructible universe $L$, and $AC$ is a theorem of the constructible universe. We have</p>\n\n<p>$$ ZF+(V=L) \\vdash AC $$</p>\n\n<p>In other words, a necessary condition for asserting $\\neg AC$ is to first assert \"there exists a set that cannot be constructed\" &mdash; that is, it requires positing the existence of additional structure above and beyond what is guaranteed by ZF.</p>\n\n<p>Thus, it seems clear to me that ZFC is by far the better choice for <em>foundations</em>. One may still wish to work with another set theoretic universe, but that's most appropriately done as the study of <em>additional structure</em> built atop of the foundations, not by rewriting the foundations themselves.</p>\n" }, { "answer_id": 260141, "score": 8, "text": "<p>Physical applications of the Banach–Tarski theorem were explored by <a href=\"https://archive.org/details/Startling_Stories_v18n03_1949-01\" rel=\"noreferrer\">Henry Kuttner, The Time Axis, <em>Startling Stories</em> 18:3 (January 1949), 13–82</a>. Some excerpts from <a href=\"https://archive.org/stream/Startling_Stories_v18n03_1949-01#page/n65/mode/2up\" rel=\"noreferrer\">pp. 66–67</a>:</p>\n\n<blockquote>\n <p>\"'Professor Raphael M. Robinson of the University of California now shows that it is possible to divide a solid sphere into a minimum of five pieces and reassemble them to form two spheres of the same size as the original one. Two of the pieces are used to form one of the new spheres and three to form the other.<br>\n <br>\"'Some of the pieces must necessarily be of such complicated structure that it is impossible to assign volume to them. Otherwise the sum of the volumes of the five pieces would have to be equal both to the volume of the original sphere and to the sum of the volumes of the two new spheres, which is twice as great.'\"<br>\n <br>[. . . .]<br>\n <br>\"This is it,\" he said.<br>\n <br>Even the crowd around the neural-web table thinned as the workers in the laboratory flocked around him to watch.<br>\n <br>He had a sphere about the size of a grapefruit, floating in mid-air above his table. He did things to it with quick flashes of light that acted exactly like knives, in that it fell apart wherever the lights touched, but I got the impression that those divisions were much less simple than knife-cuts would be. The light shivered as it slashed and the cuts must have been very complex, dividing molecules with a selective precision beyond my powers of comprehension.<br>\n <br>The sphere floated apart. It changed shape under the lights. I am pretty sure it changed shape in four dimensions, because after a while I literally could not watch any more. The shape did agonizing things to my eyes when I tried to focus on it.<br>\n <br>When I heard a long sigh go up simultaneously from the watchers I risked a look again.<br>\n <br>There were two spheres floating where one had floated before.</p>\n</blockquote>\n" }, { "answer_id": 260680, "score": 5, "text": "<p>I think it is interesting to re-phrase this question relative to <strong>other</strong> axioms and/or theorems of ZFC. What Andrej Bauer's answer suggests is that it may not be the axiom of choice <em>per se</em> that is the culprit, but rather the underlying structure.</p>\n\n<p>For example, it is provable that the existence of non-Lebesgue measurable sets and the Banach–Tarski paradox are <strong>both</strong> an implication over ZF (without Choice) of the Hahn–Banach theorem (HB) in functional analysis. This means that we can prove those \"pathologies\" as a theorems from ZF+HB. </p>\n\n<p>Another way to put it is to analyze the structure within the context of different orders of logic. In ordinary first-order logic Choice is provably equivalent to the well-ordering theorem (WO) over ZF. What is different in second-order logic is that WO is strictly stronger than choice: WO $\\vdash_{ZF}$ Choice, but Choice $\\nvdash_{ZF}$ WO.</p>\n\n<p>In other words, one may get the feeling that perhaps there has been too much emphasis put on Choice. Other axioms and theorems can prove just as problematic. The approach described above is the essence of Reverse Mathematics (RM). While RM has been traditionally carried out at a much lower proof-theoretic strength level (subsystems of second-order arithmetic), in my opinion this provides a very useful framework for analyzing the foundations of other parts of mathematics.</p>\n" }, { "answer_id": 260072, "score": 120, "text": "<p>There are <em>two</em> ingredients in the Banach-Tarski decomposition theorem:</p>\n<ol>\n<li>The notion of <em>space</em>, together with derived notions of <em>part</em> and <em>decomposition</em>.</li>\n<li>The axiom of choice.</li>\n</ol>\n<p>Most discussion about the theorem revolve around the axiom of choice. I would like to point out that the notion of space can be put under scrutiny as well.</p>\n<p>The Banach-Tarski decomposition of the sphere produces non-measurable parts of the sphere. If we <em>restrict</em> the notion of &quot;part&quot; to &quot;measurable subset&quot; the theorem disappears. For instance, if we move over into a model of set theory (without choice) in which all sets are measurable, we will have no Banach-Tarski. This is all well known.</p>\n<p>Somewhat amazingly, we can make the Banach-Tarski decomposition go away by <em>extending</em> the notion of subspace, and keep choice too. Alex Simpson in <a href=\"https://web.archive.org/web/20160909042100/http://homepages.inf.ed.ac.uk/als/Research/Sources/mrs.pdf\" rel=\"noreferrer\">Measure, Randomness and Sublocales</a> (Annals of Pure and Applied Logic, 163(11), pp. 1642-1659, 2012) shows that this is achieved by generalizing the notion of topological space to that of <a href=\"https://ncatlab.org/nlab/show/locale\" rel=\"noreferrer\">locale</a>. He explains it thus:</p>\n<blockquote>\n<p>&quot;The different pieces in the partitions defined by Vitali and by Banach and Tarski are deeply intertangled with each other. According to our notion of “part”, two such intertangled pieces are not disjoint from each other, so additivity does not apply. An intuitive explanation for the failure of disjointness is that, although two such pieces share no point in common, they nevertheless overlap on the topological “glue” that bonds neighbouring points in <span class=\"math-container\">$\\mathbb{R}^n$</span> together.&quot;</p>\n</blockquote>\n<p>Peter Johnstone explained in <a href=\"http://pointlesstopology.com/the-point-of-pointless-topology.pdf\" rel=\"noreferrer\">The point of pointless topology</a> why locales have mathematical significance that goes far beyond fixing a strange theorem about decomposition of the sphere. Why isn't everyone using locales? I do not know, I think it is purely a historic accident. At some point in the 20th century mathematicians collectively lost the idea that there is more to space than just its points.</p>\n<p>I personally prefer to blame the trouble on the notion of space, rather than the axiom of choice. As far as possible, geometric problems should be the business of geometry, not logic or set theory. Mathematicians are used to operating with various kinds of spaces (in geometry, in analysis, in topology, in algebraic geometry, in computation, etc.) and so it seems only natural that one should worry about using the correct notion of space first, and about underlying foundational principles later. Good math is immune to changes in foundations.</p>\n" }, { "answer_id": 260065, "score": 42, "text": "<p>It's notable that most of the \"bread and butter\" mathematical consequences of the axiom of choice are actually consequences of <em>countable</em> choice. (Every infinite set contains a countable subset, a countable union of countable sets is countable, etc.) The Hahn-Banach theorem is a counterexample, but only if you want it for nonseparable spaces, and I can't think of any time I've ever needed this. When restricted to separable Banach spaces it doesn't require any choice principle at all! Whereas the seemingly pathological consequences of choice (existence of nonmeasurable sets, Banach-Tarski, well-ordering of the real line) generally do not follow from countable choice.</p>\n\n<p>So the argument from mathematical value seems to me to support countable choice more than full choice. But that isn't a very strong argument, is it? We can't decide whether an axiom is true based on whether we like its consequences. At best it's suggestive.</p>\n\n<p>Incidentally, I had the impression when I read Zermelo that he had great polemical skill, but none of his arguments seemed to get directly to the <em>truth</em> of the axiom. He argues for the mathematical value of the axiom. He points out that his critics have themselves on occasion unwittingly used the axiom, which is a devastating point, but has little bearing on the question of truth. (If I'm not mistaken, those unwitting uses were all of countable choice, by the way.)</p>\n\n<p>You ask if one should \"believe\" the axiom of choice, and I think you are right to put the word \"believe\" in quotes. I feel strongly that set-theoretic assertions are objectively meaningful, but I also feel that philosophers of mathematics have done a very poor job of clarifying what sets are. (Halmos: \"A pack of wolves, a bunch of grapes, or a flock of pigeons are all examples of sets of things.\" Black: \"It ought then to make sense, at least sometimes, to speak of being pursued by a set, or eating a set, or putting a set to flight.\") If we can't even get that straight, it's hard to come to grips with questions about the truth of questionable axioms.</p>\n" }, { "answer_id": 260120, "score": 36, "text": "<p>The other answers don't seem to have said much about why the axiom of choice is widely regarded as plausible. Let me try to address that question.</p>\n\n<p>First let's dispose of some non-reasons. In response to your questions, I don't know of anyone who thinks that the Banach&ndash;Tarski paradox is a reason to believe in the axiom of choice. I also don't know of anyone who argues, \"It is a priori plausible that there exist non-measurable sets; so the fact that the axiom of choice yields the attractive conclusion that there are non-measurable sets is a point in favor of believing the axiom of choice.\" Instead, those who are comfortable with the existence of non-measurable sets typically start by accepting the axiom of choice, and then they accept non-measurable sets as \"part of the territory\" that comes with the axiom of choice.</p>\n\n<p>Those who think that Banach&ndash;Tarski casts doubt on the axiom of choice typically have a philosophical predisposition that <i>math is supposed to model the physical world closely</i>. So for example, $\\mathbb R^3$ is not just a random mathematical structure that we study purely for its own sake; it is supposed to be a decent model of physical space (or at least, open subsets of $\\mathbb R^3$ are supposed to model localized regions of physical space). Banach&ndash;Tarski, when given a direct physical interpretation in this way, yields something that we \"know\" makes no physical sense, and so if we think that math is supposed to yield physical truth in this way, then Banach&ndash;Tarski is going to lead us to reject something in the math. Whether that \"something\" we reject is the axiom of choice is a separate question, and Andrej Bauer's excellent answer shows that there are other options, but the point I want to highlight is that we're going to be led down this path in the first place only if we have certain presuppositions about how math and physics are supposed to relate.</p>\n\n<p>There are others who don't view set theory in this way. According to them, set theory is supposed to be about <i>abstract collections of things</i>, and the way to arrive at axioms is by abstractly thinking about what properties they should have, not by comparing them with the physical world. The axiom of choice can be thought of as saying that if you have a bunch of nonempty collections of things, then there is another collection of things that contains one element from each of your original nonempty collections. Stated this way, the principle sounds intuitively plausible, and I would argue that this intuitive plausibility is, at least implicitly, the main argument in the minds of most people who accept the axiom of choice. If this is the way you think, then non-measurable sets and Banach&ndash;Tarski are not going to dissuade you from accepting the axiom of choice. Those phenomena will just lead you to say that we can't arrive at physical predictions from mathematics in such a naive manner; instead, to do physics, we have to formulate <i>physical theories</i>. Math can of course help a lot with the construction of physical theories, but it's not as simple as just saying that the mathematical theory of $\\mathbb R^3$ <i>is</i> our theory of physical space.</p>\n\n<p>These two options aren't the only options. The <a href=\"https://en.wikipedia.org/wiki/Solovay_model\">work of Solovay</a> shows that you can, to a large extent, <a href=\"https://books.google.com/ngrams/graph?content=have+your+cake+and+eat%2Ceat+your+cake+and+have&amp;year_start=1840&amp;year_end=2000&amp;corpus=15&amp;smoothing=3&amp;share=&amp;direct_url=t1%3B%2Chave%20your%20cake%20and%20eat%3B%2Cc0%3B.t1%3B%2Ceat%20your%20cake%20and%20have%3B%2Cc0\">have your cake and eat it too</a>, by working in a set-theoretic universe where all subsets of $\\mathbb R^n$ are Lebesgue-measurable and a weakened, but still quite strong, version of the axiom of choice known as \"dependent choice\" is available. Why Solovay's model hasn't become more popular is not completely clear, but perhaps part of the reason is that it feels like a \"compromise position,\" and the people in the two different camps above have not seen any need to migrate to that kind of compromise.</p>\n" }, { "answer_id": 260118, "score": 27, "text": "<blockquote>\n <p>Are there reasons why it is plausible (for physicists, philosophers,\n mathematicians) to believe that not all sets should be measurable?</p>\n</blockquote>\n\n<p>Yes. If every set of reals is Lebesgue measurable, then you can partition $\\mathbb{R}$ into <em>more than continuum many</em> pairwise disjoint non-empty pieces. (See <a href=\"https://mathoverflow.net/a/22935/33039\">this answer and comments</a> for the details.)</p>\n\n<p>Surely the Banach-Tarski paradox seems unintuitive. But having a set that can be broken up into more pieces than there originally were... is just wrong.</p>\n" }, { "answer_id": 260092, "score": 18, "text": "<p>Arguments from physics may not help. Here is Bryce DeWitt <a href=\"http://science.sciencemag.org/content/182/4113/705\">reviewing</a> Stephen Hawking and G.F.R. Ellis using the axiom of choice in 1973:</p>\n\n<blockquote>\n <p>The book also contains one failure to distinguish between mathematics\n and physics that is actually serious. This is in the proof of the\n main theorem of chapter 7, that given a set of Cauchy data on a smooth\n spacelike hypersurface there exists a unique maximal development\n therefrom of Einstein’s empty-space equations. The proof, essentially\n due to Choquet-Bruhat and Geroch, makes use of the axiom of choice,\n in the guise of Zorn’s lemma. Now mathematicians may use this axiom if\n they wish, but it has no place in physics. Physicists are already\n stretching things, from an operational standpoint, in using the axiom\n of infinity.</p>\n \n <p>It is not a question here of resurrecting an old and out-of-date\n mathematical controversy. The simple fact is that the axiom of\n choice never is really needed except when dealing with sets and\n relations in non-constructive ways. Many remarkable and beautiful\n theorems can be proved only with its aid. But its irrelevance to\n physics should be evident from the fact that its denial, as Paul\n Cohen has shown us, is equally consistent with the other axioms\n of set theory. And these other axioms suffice for the constructions of\n the real numbers, Hilbert spaces, C* algebras, and pseudo-Riemannian\n manifolds–that is, of all the paraphernalia of theoretical physics.</p>\n \n <p>In “proving” the global Cauchy development theorem with the aid of\n Zorn’s lemma what one is actually doing is assuming that a “choice\n function” exists for every set of developments extending a given\n Cauchy development. This, of course, is begging the question. The\n physicist’s job is not done until he can show, by an explicit\n algorithm or construction, how one could in principle always select a\n member from every such set of developments. Failing that he has proved\n nothing.</p>\n</blockquote>\n\n<p>Some physicists want to use the axiom of choice, but some physicists don't.</p>\n" }, { "answer_id": 260073, "score": 13, "text": "<p>Just a quick addendum: the result was actually found by Felix Hausdorff and then repackaged in a more spectacular form by Banach and Tarski. Hausdorff' point was precisely to show that the axiom of choice leads to such unreasonable consequences that it should probably be avoided. This was a hot topic of discussion during the following decades, and it seems that the answer at large from the mathematical community is: ok, there are inconveniences, but the advantages of using the AC are superior and we prefer to have it available.</p>\n" }, { "answer_id": 260128, "score": 9, "text": "<p>I think one of the strongest arguments for the axiom of choice is that every model of ZF contains as an inner model a constructible universe $L$, and $AC$ is a theorem of the constructible universe. We have</p>\n\n<p>$$ ZF+(V=L) \\vdash AC $$</p>\n\n<p>In other words, a necessary condition for asserting $\\neg AC$ is to first assert \"there exists a set that cannot be constructed\" &mdash; that is, it requires positing the existence of additional structure above and beyond what is guaranteed by ZF.</p>\n\n<p>Thus, it seems clear to me that ZFC is by far the better choice for <em>foundations</em>. One may still wish to work with another set theoretic universe, but that's most appropriately done as the study of <em>additional structure</em> built atop of the foundations, not by rewriting the foundations themselves.</p>\n" }, { "answer_id": 260141, "score": 8, "text": "<p>Physical applications of the Banach–Tarski theorem were explored by <a href=\"https://archive.org/details/Startling_Stories_v18n03_1949-01\" rel=\"noreferrer\">Henry Kuttner, The Time Axis, <em>Startling Stories</em> 18:3 (January 1949), 13–82</a>. Some excerpts from <a href=\"https://archive.org/stream/Startling_Stories_v18n03_1949-01#page/n65/mode/2up\" rel=\"noreferrer\">pp. 66–67</a>:</p>\n\n<blockquote>\n <p>\"'Professor Raphael M. Robinson of the University of California now shows that it is possible to divide a solid sphere into a minimum of five pieces and reassemble them to form two spheres of the same size as the original one. Two of the pieces are used to form one of the new spheres and three to form the other.<br>\n <br>\"'Some of the pieces must necessarily be of such complicated structure that it is impossible to assign volume to them. Otherwise the sum of the volumes of the five pieces would have to be equal both to the volume of the original sphere and to the sum of the volumes of the two new spheres, which is twice as great.'\"<br>\n <br>[. . . .]<br>\n <br>\"This is it,\" he said.<br>\n <br>Even the crowd around the neural-web table thinned as the workers in the laboratory flocked around him to watch.<br>\n <br>He had a sphere about the size of a grapefruit, floating in mid-air above his table. He did things to it with quick flashes of light that acted exactly like knives, in that it fell apart wherever the lights touched, but I got the impression that those divisions were much less simple than knife-cuts would be. The light shivered as it slashed and the cuts must have been very complex, dividing molecules with a selective precision beyond my powers of comprehension.<br>\n <br>The sphere floated apart. It changed shape under the lights. I am pretty sure it changed shape in four dimensions, because after a while I literally could not watch any more. The shape did agonizing things to my eyes when I tried to focus on it.<br>\n <br>When I heard a long sigh go up simultaneously from the watchers I risked a look again.<br>\n <br>There were two spheres floating where one had floated before.</p>\n</blockquote>\n" }, { "answer_id": 260680, "score": 5, "text": "<p>I think it is interesting to re-phrase this question relative to <strong>other</strong> axioms and/or theorems of ZFC. What Andrej Bauer's answer suggests is that it may not be the axiom of choice <em>per se</em> that is the culprit, but rather the underlying structure.</p>\n\n<p>For example, it is provable that the existence of non-Lebesgue measurable sets and the Banach–Tarski paradox are <strong>both</strong> an implication over ZF (without Choice) of the Hahn–Banach theorem (HB) in functional analysis. This means that we can prove those \"pathologies\" as a theorems from ZF+HB. </p>\n\n<p>Another way to put it is to analyze the structure within the context of different orders of logic. In ordinary first-order logic Choice is provably equivalent to the well-ordering theorem (WO) over ZF. What is different in second-order logic is that WO is strictly stronger than choice: WO $\\vdash_{ZF}$ Choice, but Choice $\\nvdash_{ZF}$ WO.</p>\n\n<p>In other words, one may get the feeling that perhaps there has been too much emphasis put on Choice. Other axioms and theorems can prove just as problematic. The approach described above is the essence of Reverse Mathematics (RM). While RM has been traditionally carried out at a much lower proof-theoretic strength level (subsystems of second-order arithmetic), in my opinion this provides a very useful framework for analyzing the foundations of other parts of mathematics.</p>\n" } ]
Question: Axiom of choice, Banach-Tarski and reality Answer 1: <p>There are <em>two</em> ingredients in the Banach-Tarski decomposition theorem:</p> <ol> <li>The notion of <em>space</em>, together with derived notions of <em>part</em> and <em>decomposition</em>.</li> <li>The axiom of choice.</li> </ol> <p>Most discussion about the theorem revolve around the axiom of choice. I would like to point out that the notion of space can be put under scrutiny as well.</p> <p>The Banach-Tarski decomposition of the sphere produces non-measurable parts of the sphere. If we <em>restrict</em> the notion of &quot;part&quot; to &quot;measurable subset&quot; the theorem disappears. For instance, if we move over into a model of set theory (without choice) in which all sets are measurable, we will have no Banach-Tarski. This is all well known.</p> <p>Somewhat amazingly, we can make the Banach-Tarski decomposition go away by <em>extending</em> the notion of subspace, and keep choice too. Alex Simpson in <a href="https://web.archive.org/web/20160909042100/http://homepages.inf.ed.ac.uk/als/Research/Sources/mrs.pdf" rel="noreferrer">Measure, Randomness and Sublocales</a> (Annals of Pure and Applied Logic, 163(11), pp. 1642-1659, 2012) shows that this is achieved by generalizing the notion of topological space to that of <a href="https://ncatlab.org/nlab/show/locale" rel="noreferrer">locale</a>. He explains it thus:</p> <blockquote> <p>&quot;The different pieces in the partitions defined by Vitali and by Banach and Tarski are deeply intertangled with each other. According to our notion of “part”, two such intertangled pieces are not disjoint from each other, so additivity does not apply. An intuitive explanation for the failure of disjointness is that, although two such pieces share no point in common, they nevertheless overlap on the topological “glue” that bonds neighbouring points in <span class="math-container">$\mathbb{R}^n$</span> together.&quot;</p> </blockquote> <p>Peter Johnstone explained in <a href="http://pointlesstopology.com/the-point-of-pointless-topology.pdf" rel="noreferrer">The point of pointless topology</a> why locales have mathematical significance that goes far beyond fixing a strange theorem about decomposition of the sphere. Why isn't everyone using locales? I do not know, I think it is purely a historic accident. At some point in the 20th century mathematicians collectively lost the idea that there is more to space than just its points.</p> <p>I personally prefer to blame the trouble on the notion of space, rather than the axiom of choice. As far as possible, geometric problems should be the business of geometry, not logic or set theory. Mathematicians are used to operating with various kinds of spaces (in geometry, in analysis, in topology, in algebraic geometry, in computation, etc.) and so it seems only natural that one should worry about using the correct notion of space first, and about underlying foundational principles later. Good math is immune to changes in foundations.</p> Answer 2: <p>It's notable that most of the "bread and butter" mathematical consequences of the axiom of choice are actually consequences of <em>countable</em> choice. (Every infinite set contains a countable subset, a countable union of countable sets is countable, etc.) The Hahn-Banach theorem is a counterexample, but only if you want it for nonseparable spaces, and I can't think of any time I've ever needed this. When restricted to separable Banach spaces it doesn't require any choice principle at all! Whereas the seemingly pathological consequences of choice (existence of nonmeasurable sets, Banach-Tarski, well-ordering of the real line) generally do not follow from countable choice.</p> <p>So the argument from mathematical value seems to me to support countable choice more than full choice. But that isn't a very strong argument, is it? We can't decide whether an axiom is true based on whether we like its consequences. At best it's suggestive.</p> <p>Incidentally, I had the impression when I read Zermelo that he had great polemical skill, but none of his arguments seemed to get directly to the <em>truth</em> of the axiom. He argues for the mathematical value of the axiom. He points out that his critics have themselves on occasion unwittingly used the axiom, which is a devastating point, but has little bearing on the question of truth. (If I'm not mistaken, those unwitting uses were all of countable choice, by the way.)</p> <p>You ask if one should "believe" the axiom of choice, and I think you are right to put the word "believe" in quotes. I feel strongly that set-theoretic assertions are objectively meaningful, but I also feel that philosophers of mathematics have done a very poor job of clarifying what sets are. (Halmos: "A pack of wolves, a bunch of grapes, or a flock of pigeons are all examples of sets of things." Black: "It ought then to make sense, at least sometimes, to speak of being pursued by a set, or eating a set, or putting a set to flight.") If we can't even get that straight, it's hard to come to grips with questions about the truth of questionable axioms.</p> Answer 3: <p>The other answers don't seem to have said much about why the axiom of choice is widely regarded as plausible. Let me try to address that question.</p> <p>First let's dispose of some non-reasons. In response to your questions, I don't know of anyone who thinks that the Banach&ndash;Tarski paradox is a reason to believe in the axiom of choice. I also don't know of anyone who argues, "It is a priori plausible that there exist non-measurable sets; so the fact that the axiom of choice yields the attractive conclusion that there are non-measurable sets is a point in favor of believing the axiom of choice." Instead, those who are comfortable with the existence of non-measurable sets typically start by accepting the axiom of choice, and then they accept non-measurable sets as "part of the territory" that comes with the axiom of choice.</p> <p>Those who think that Banach&ndash;Tarski casts doubt on the axiom of choice typically have a philosophical predisposition that <i>math is supposed to model the physical world closely</i>. So for example, $\mathbb R^3$ is not just a random mathematical structure that we study purely for its own sake; it is supposed to be a decent model of physical space (or at least, open subsets of $\mathbb R^3$ are supposed to model localized regions of physical space). Banach&ndash;Tarski, when given a direct physical interpretation in this way, yields something that we "know" makes no physical sense, and so if we think that math is supposed to yield physical truth in this way, then Banach&ndash;Tarski is going to lead us to reject something in the math. Whether that "something" we reject is the axiom of choice is a separate question, and Andrej Bauer's excellent answer shows that there are other options, but the point I want to highlight is that we're going to be led down this path in the first place only if we have certain presuppositions about how math and physics are supposed to relate.</p> <p>There are others who don't view set theory in this way. According to them, set theory is supposed to be about <i>abstract collections of things</i>, and the way to arrive at axioms is by abstractly thinking about what properties they should have, not by comparing them with the physical world. The axiom of choice can be thought of as saying that if you have a bunch of nonempty collections of things, then there is another collection of things that contains one element from each of your original nonempty collections. Stated this way, the principle sounds intuitively plausible, and I would argue that this intuitive plausibility is, at least implicitly, the main argument in the minds of most people who accept the axiom of choice. If this is the way you think, then non-measurable sets and Banach&ndash;Tarski are not going to dissuade you from accepting the axiom of choice. Those phenomena will just lead you to say that we can't arrive at physical predictions from mathematics in such a naive manner; instead, to do physics, we have to formulate <i>physical theories</i>. Math can of course help a lot with the construction of physical theories, but it's not as simple as just saying that the mathematical theory of $\mathbb R^3$ <i>is</i> our theory of physical space.</p> <p>These two options aren't the only options. The <a href="https://en.wikipedia.org/wiki/Solovay_model">work of Solovay</a> shows that you can, to a large extent, <a href="https://books.google.com/ngrams/graph?content=have+your+cake+and+eat%2Ceat+your+cake+and+have&amp;year_start=1840&amp;year_end=2000&amp;corpus=15&amp;smoothing=3&amp;share=&amp;direct_url=t1%3B%2Chave%20your%20cake%20and%20eat%3B%2Cc0%3B.t1%3B%2Ceat%20your%20cake%20and%20have%3B%2Cc0">have your cake and eat it too</a>, by working in a set-theoretic universe where all subsets of $\mathbb R^n$ are Lebesgue-measurable and a weakened, but still quite strong, version of the axiom of choice known as "dependent choice" is available. Why Solovay's model hasn't become more popular is not completely clear, but perhaps part of the reason is that it feels like a "compromise position," and the people in the two different camps above have not seen any need to migrate to that kind of compromise.</p> Answer 4: <blockquote> <p>Are there reasons why it is plausible (for physicists, philosophers, mathematicians) to believe that not all sets should be measurable?</p> </blockquote> <p>Yes. If every set of reals is Lebesgue measurable, then you can partition $\mathbb{R}$ into <em>more than continuum many</em> pairwise disjoint non-empty pieces. (See <a href="https://mathoverflow.net/a/22935/33039">this answer and comments</a> for the details.)</p> <p>Surely the Banach-Tarski paradox seems unintuitive. But having a set that can be broken up into more pieces than there originally were... is just wrong.</p> Answer 5: <p>Arguments from physics may not help. Here is Bryce DeWitt <a href="http://science.sciencemag.org/content/182/4113/705">reviewing</a> Stephen Hawking and G.F.R. Ellis using the axiom of choice in 1973:</p> <blockquote> <p>The book also contains one failure to distinguish between mathematics and physics that is actually serious. This is in the proof of the main theorem of chapter 7, that given a set of Cauchy data on a smooth spacelike hypersurface there exists a unique maximal development therefrom of Einstein’s empty-space equations. The proof, essentially due to Choquet-Bruhat and Geroch, makes use of the axiom of choice, in the guise of Zorn’s lemma. Now mathematicians may use this axiom if they wish, but it has no place in physics. Physicists are already stretching things, from an operational standpoint, in using the axiom of infinity.</p> <p>It is not a question here of resurrecting an old and out-of-date mathematical controversy. The simple fact is that the axiom of choice never is really needed except when dealing with sets and relations in non-constructive ways. Many remarkable and beautiful theorems can be proved only with its aid. But its irrelevance to physics should be evident from the fact that its denial, as Paul Cohen has shown us, is equally consistent with the other axioms of set theory. And these other axioms suffice for the constructions of the real numbers, Hilbert spaces, C* algebras, and pseudo-Riemannian manifolds–that is, of all the paraphernalia of theoretical physics.</p> <p>In “proving” the global Cauchy development theorem with the aid of Zorn’s lemma what one is actually doing is assuming that a “choice function” exists for every set of developments extending a given Cauchy development. This, of course, is begging the question. The physicist’s job is not done until he can show, by an explicit algorithm or construction, how one could in principle always select a member from every such set of developments. Failing that he has proved nothing.</p> </blockquote> <p>Some physicists want to use the axiom of choice, but some physicists don't.</p> Answer 6: <p>Just a quick addendum: the result was actually found by Felix Hausdorff and then repackaged in a more spectacular form by Banach and Tarski. Hausdorff' point was precisely to show that the axiom of choice leads to such unreasonable consequences that it should probably be avoided. This was a hot topic of discussion during the following decades, and it seems that the answer at large from the mathematical community is: ok, there are inconveniences, but the advantages of using the AC are superior and we prefer to have it available.</p> Answer 7: <p>I think one of the strongest arguments for the axiom of choice is that every model of ZF contains as an inner model a constructible universe $L$, and $AC$ is a theorem of the constructible universe. We have</p> <p>$$ ZF+(V=L) \vdash AC $$</p> <p>In other words, a necessary condition for asserting $\neg AC$ is to first assert "there exists a set that cannot be constructed" &mdash; that is, it requires positing the existence of additional structure above and beyond what is guaranteed by ZF.</p> <p>Thus, it seems clear to me that ZFC is by far the better choice for <em>foundations</em>. One may still wish to work with another set theoretic universe, but that's most appropriately done as the study of <em>additional structure</em> built atop of the foundations, not by rewriting the foundations themselves.</p> Answer 8: <p>Physical applications of the Banach–Tarski theorem were explored by <a href="https://archive.org/details/Startling_Stories_v18n03_1949-01" rel="noreferrer">Henry Kuttner, The Time Axis, <em>Startling Stories</em> 18:3 (January 1949), 13–82</a>. Some excerpts from <a href="https://archive.org/stream/Startling_Stories_v18n03_1949-01#page/n65/mode/2up" rel="noreferrer">pp. 66–67</a>:</p> <blockquote> <p>"'Professor Raphael M. Robinson of the University of California now shows that it is possible to divide a solid sphere into a minimum of five pieces and reassemble them to form two spheres of the same size as the original one. Two of the pieces are used to form one of the new spheres and three to form the other.<br> <br>"'Some of the pieces must necessarily be of such complicated structure that it is impossible to assign volume to them. Otherwise the sum of the volumes of the five pieces would have to be equal both to the volume of the original sphere and to the sum of the volumes of the two new spheres, which is twice as great.'"<br> <br>[. . . .]<br> <br>"This is it," he said.<br> <br>Even the crowd around the neural-web table thinned as the workers in the laboratory flocked around him to watch.<br> <br>He had a sphere about the size of a grapefruit, floating in mid-air above his table. He did things to it with quick flashes of light that acted exactly like knives, in that it fell apart wherever the lights touched, but I got the impression that those divisions were much less simple than knife-cuts would be. The light shivered as it slashed and the cuts must have been very complex, dividing molecules with a selective precision beyond my powers of comprehension.<br> <br>The sphere floated apart. It changed shape under the lights. I am pretty sure it changed shape in four dimensions, because after a while I literally could not watch any more. The shape did agonizing things to my eyes when I tried to focus on it.<br> <br>When I heard a long sigh go up simultaneously from the watchers I risked a look again.<br> <br>There were two spheres floating where one had floated before.</p> </blockquote> Answer 9: <p>I think it is interesting to re-phrase this question relative to <strong>other</strong> axioms and/or theorems of ZFC. What Andrej Bauer's answer suggests is that it may not be the axiom of choice <em>per se</em> that is the culprit, but rather the underlying structure.</p> <p>For example, it is provable that the existence of non-Lebesgue measurable sets and the Banach–Tarski paradox are <strong>both</strong> an implication over ZF (without Choice) of the Hahn–Banach theorem (HB) in functional analysis. This means that we can prove those "pathologies" as a theorems from ZF+HB. </p> <p>Another way to put it is to analyze the structure within the context of different orders of logic. In ordinary first-order logic Choice is provably equivalent to the well-ordering theorem (WO) over ZF. What is different in second-order logic is that WO is strictly stronger than choice: WO $\vdash_{ZF}$ Choice, but Choice $\nvdash_{ZF}$ WO.</p> <p>In other words, one may get the feeling that perhaps there has been too much emphasis put on Choice. Other axioms and theorems can prove just as problematic. The approach described above is the essence of Reverse Mathematics (RM). While RM has been traditionally carried out at a much lower proof-theoretic strength level (subsystems of second-order arithmetic), in my opinion this provides a very useful framework for analyzing the foundations of other parts of mathematics.</p> Answer 10: <p>There are <em>two</em> ingredients in the Banach-Tarski decomposition theorem:</p> <ol> <li>The notion of <em>space</em>, together with derived notions of <em>part</em> and <em>decomposition</em>.</li> <li>The axiom of choice.</li> </ol> <p>Most discussion about the theorem revolve around the axiom of choice. I would like to point out that the notion of space can be put under scrutiny as well.</p> <p>The Banach-Tarski decomposition of the sphere produces non-measurable parts of the sphere. If we <em>restrict</em> the notion of &quot;part&quot; to &quot;measurable subset&quot; the theorem disappears. For instance, if we move over into a model of set theory (without choice) in which all sets are measurable, we will have no Banach-Tarski. This is all well known.</p> <p>Somewhat amazingly, we can make the Banach-Tarski decomposition go away by <em>extending</em> the notion of subspace, and keep choice too. Alex Simpson in <a href="https://web.archive.org/web/20160909042100/http://homepages.inf.ed.ac.uk/als/Research/Sources/mrs.pdf" rel="noreferrer">Measure, Randomness and Sublocales</a> (Annals of Pure and Applied Logic, 163(11), pp. 1642-1659, 2012) shows that this is achieved by generalizing the notion of topological space to that of <a href="https://ncatlab.org/nlab/show/locale" rel="noreferrer">locale</a>. He explains it thus:</p> <blockquote> <p>&quot;The different pieces in the partitions defined by Vitali and by Banach and Tarski are deeply intertangled with each other. According to our notion of “part”, two such intertangled pieces are not disjoint from each other, so additivity does not apply. An intuitive explanation for the failure of disjointness is that, although two such pieces share no point in common, they nevertheless overlap on the topological “glue” that bonds neighbouring points in <span class="math-container">$\mathbb{R}^n$</span> together.&quot;</p> </blockquote> <p>Peter Johnstone explained in <a href="http://pointlesstopology.com/the-point-of-pointless-topology.pdf" rel="noreferrer">The point of pointless topology</a> why locales have mathematical significance that goes far beyond fixing a strange theorem about decomposition of the sphere. Why isn't everyone using locales? I do not know, I think it is purely a historic accident. At some point in the 20th century mathematicians collectively lost the idea that there is more to space than just its points.</p> <p>I personally prefer to blame the trouble on the notion of space, rather than the axiom of choice. As far as possible, geometric problems should be the business of geometry, not logic or set theory. Mathematicians are used to operating with various kinds of spaces (in geometry, in analysis, in topology, in algebraic geometry, in computation, etc.) and so it seems only natural that one should worry about using the correct notion of space first, and about underlying foundational principles later. Good math is immune to changes in foundations.</p> Answer 11: <p>It's notable that most of the "bread and butter" mathematical consequences of the axiom of choice are actually consequences of <em>countable</em> choice. (Every infinite set contains a countable subset, a countable union of countable sets is countable, etc.) The Hahn-Banach theorem is a counterexample, but only if you want it for nonseparable spaces, and I can't think of any time I've ever needed this. When restricted to separable Banach spaces it doesn't require any choice principle at all! Whereas the seemingly pathological consequences of choice (existence of nonmeasurable sets, Banach-Tarski, well-ordering of the real line) generally do not follow from countable choice.</p> <p>So the argument from mathematical value seems to me to support countable choice more than full choice. But that isn't a very strong argument, is it? We can't decide whether an axiom is true based on whether we like its consequences. At best it's suggestive.</p> <p>Incidentally, I had the impression when I read Zermelo that he had great polemical skill, but none of his arguments seemed to get directly to the <em>truth</em> of the axiom. He argues for the mathematical value of the axiom. He points out that his critics have themselves on occasion unwittingly used the axiom, which is a devastating point, but has little bearing on the question of truth. (If I'm not mistaken, those unwitting uses were all of countable choice, by the way.)</p> <p>You ask if one should "believe" the axiom of choice, and I think you are right to put the word "believe" in quotes. I feel strongly that set-theoretic assertions are objectively meaningful, but I also feel that philosophers of mathematics have done a very poor job of clarifying what sets are. (Halmos: "A pack of wolves, a bunch of grapes, or a flock of pigeons are all examples of sets of things." Black: "It ought then to make sense, at least sometimes, to speak of being pursued by a set, or eating a set, or putting a set to flight.") If we can't even get that straight, it's hard to come to grips with questions about the truth of questionable axioms.</p> Answer 12: <p>The other answers don't seem to have said much about why the axiom of choice is widely regarded as plausible. Let me try to address that question.</p> <p>First let's dispose of some non-reasons. In response to your questions, I don't know of anyone who thinks that the Banach&ndash;Tarski paradox is a reason to believe in the axiom of choice. I also don't know of anyone who argues, "It is a priori plausible that there exist non-measurable sets; so the fact that the axiom of choice yields the attractive conclusion that there are non-measurable sets is a point in favor of believing the axiom of choice." Instead, those who are comfortable with the existence of non-measurable sets typically start by accepting the axiom of choice, and then they accept non-measurable sets as "part of the territory" that comes with the axiom of choice.</p> <p>Those who think that Banach&ndash;Tarski casts doubt on the axiom of choice typically have a philosophical predisposition that <i>math is supposed to model the physical world closely</i>. So for example, $\mathbb R^3$ is not just a random mathematical structure that we study purely for its own sake; it is supposed to be a decent model of physical space (or at least, open subsets of $\mathbb R^3$ are supposed to model localized regions of physical space). Banach&ndash;Tarski, when given a direct physical interpretation in this way, yields something that we "know" makes no physical sense, and so if we think that math is supposed to yield physical truth in this way, then Banach&ndash;Tarski is going to lead us to reject something in the math. Whether that "something" we reject is the axiom of choice is a separate question, and Andrej Bauer's excellent answer shows that there are other options, but the point I want to highlight is that we're going to be led down this path in the first place only if we have certain presuppositions about how math and physics are supposed to relate.</p> <p>There are others who don't view set theory in this way. According to them, set theory is supposed to be about <i>abstract collections of things</i>, and the way to arrive at axioms is by abstractly thinking about what properties they should have, not by comparing them with the physical world. The axiom of choice can be thought of as saying that if you have a bunch of nonempty collections of things, then there is another collection of things that contains one element from each of your original nonempty collections. Stated this way, the principle sounds intuitively plausible, and I would argue that this intuitive plausibility is, at least implicitly, the main argument in the minds of most people who accept the axiom of choice. If this is the way you think, then non-measurable sets and Banach&ndash;Tarski are not going to dissuade you from accepting the axiom of choice. Those phenomena will just lead you to say that we can't arrive at physical predictions from mathematics in such a naive manner; instead, to do physics, we have to formulate <i>physical theories</i>. Math can of course help a lot with the construction of physical theories, but it's not as simple as just saying that the mathematical theory of $\mathbb R^3$ <i>is</i> our theory of physical space.</p> <p>These two options aren't the only options. The <a href="https://en.wikipedia.org/wiki/Solovay_model">work of Solovay</a> shows that you can, to a large extent, <a href="https://books.google.com/ngrams/graph?content=have+your+cake+and+eat%2Ceat+your+cake+and+have&amp;year_start=1840&amp;year_end=2000&amp;corpus=15&amp;smoothing=3&amp;share=&amp;direct_url=t1%3B%2Chave%20your%20cake%20and%20eat%3B%2Cc0%3B.t1%3B%2Ceat%20your%20cake%20and%20have%3B%2Cc0">have your cake and eat it too</a>, by working in a set-theoretic universe where all subsets of $\mathbb R^n$ are Lebesgue-measurable and a weakened, but still quite strong, version of the axiom of choice known as "dependent choice" is available. Why Solovay's model hasn't become more popular is not completely clear, but perhaps part of the reason is that it feels like a "compromise position," and the people in the two different camps above have not seen any need to migrate to that kind of compromise.</p> Answer 13: <blockquote> <p>Are there reasons why it is plausible (for physicists, philosophers, mathematicians) to believe that not all sets should be measurable?</p> </blockquote> <p>Yes. If every set of reals is Lebesgue measurable, then you can partition $\mathbb{R}$ into <em>more than continuum many</em> pairwise disjoint non-empty pieces. (See <a href="https://mathoverflow.net/a/22935/33039">this answer and comments</a> for the details.)</p> <p>Surely the Banach-Tarski paradox seems unintuitive. But having a set that can be broken up into more pieces than there originally were... is just wrong.</p> Answer 14: <p>Arguments from physics may not help. Here is Bryce DeWitt <a href="http://science.sciencemag.org/content/182/4113/705">reviewing</a> Stephen Hawking and G.F.R. Ellis using the axiom of choice in 1973:</p> <blockquote> <p>The book also contains one failure to distinguish between mathematics and physics that is actually serious. This is in the proof of the main theorem of chapter 7, that given a set of Cauchy data on a smooth spacelike hypersurface there exists a unique maximal development therefrom of Einstein’s empty-space equations. The proof, essentially due to Choquet-Bruhat and Geroch, makes use of the axiom of choice, in the guise of Zorn’s lemma. Now mathematicians may use this axiom if they wish, but it has no place in physics. Physicists are already stretching things, from an operational standpoint, in using the axiom of infinity.</p> <p>It is not a question here of resurrecting an old and out-of-date mathematical controversy. The simple fact is that the axiom of choice never is really needed except when dealing with sets and relations in non-constructive ways. Many remarkable and beautiful theorems can be proved only with its aid. But its irrelevance to physics should be evident from the fact that its denial, as Paul Cohen has shown us, is equally consistent with the other axioms of set theory. And these other axioms suffice for the constructions of the real numbers, Hilbert spaces, C* algebras, and pseudo-Riemannian manifolds–that is, of all the paraphernalia of theoretical physics.</p> <p>In “proving” the global Cauchy development theorem with the aid of Zorn’s lemma what one is actually doing is assuming that a “choice function” exists for every set of developments extending a given Cauchy development. This, of course, is begging the question. The physicist’s job is not done until he can show, by an explicit algorithm or construction, how one could in principle always select a member from every such set of developments. Failing that he has proved nothing.</p> </blockquote> <p>Some physicists want to use the axiom of choice, but some physicists don't.</p> Answer 15: <p>Just a quick addendum: the result was actually found by Felix Hausdorff and then repackaged in a more spectacular form by Banach and Tarski. Hausdorff' point was precisely to show that the axiom of choice leads to such unreasonable consequences that it should probably be avoided. This was a hot topic of discussion during the following decades, and it seems that the answer at large from the mathematical community is: ok, there are inconveniences, but the advantages of using the AC are superior and we prefer to have it available.</p> Answer 16: <p>I think one of the strongest arguments for the axiom of choice is that every model of ZF contains as an inner model a constructible universe $L$, and $AC$ is a theorem of the constructible universe. We have</p> <p>$$ ZF+(V=L) \vdash AC $$</p> <p>In other words, a necessary condition for asserting $\neg AC$ is to first assert "there exists a set that cannot be constructed" &mdash; that is, it requires positing the existence of additional structure above and beyond what is guaranteed by ZF.</p> <p>Thus, it seems clear to me that ZFC is by far the better choice for <em>foundations</em>. One may still wish to work with another set theoretic universe, but that's most appropriately done as the study of <em>additional structure</em> built atop of the foundations, not by rewriting the foundations themselves.</p> Answer 17: <p>Physical applications of the Banach–Tarski theorem were explored by <a href="https://archive.org/details/Startling_Stories_v18n03_1949-01" rel="noreferrer">Henry Kuttner, The Time Axis, <em>Startling Stories</em> 18:3 (January 1949), 13–82</a>. Some excerpts from <a href="https://archive.org/stream/Startling_Stories_v18n03_1949-01#page/n65/mode/2up" rel="noreferrer">pp. 66–67</a>:</p> <blockquote> <p>"'Professor Raphael M. Robinson of the University of California now shows that it is possible to divide a solid sphere into a minimum of five pieces and reassemble them to form two spheres of the same size as the original one. Two of the pieces are used to form one of the new spheres and three to form the other.<br> <br>"'Some of the pieces must necessarily be of such complicated structure that it is impossible to assign volume to them. Otherwise the sum of the volumes of the five pieces would have to be equal both to the volume of the original sphere and to the sum of the volumes of the two new spheres, which is twice as great.'"<br> <br>[. . . .]<br> <br>"This is it," he said.<br> <br>Even the crowd around the neural-web table thinned as the workers in the laboratory flocked around him to watch.<br> <br>He had a sphere about the size of a grapefruit, floating in mid-air above his table. He did things to it with quick flashes of light that acted exactly like knives, in that it fell apart wherever the lights touched, but I got the impression that those divisions were much less simple than knife-cuts would be. The light shivered as it slashed and the cuts must have been very complex, dividing molecules with a selective precision beyond my powers of comprehension.<br> <br>The sphere floated apart. It changed shape under the lights. I am pretty sure it changed shape in four dimensions, because after a while I literally could not watch any more. The shape did agonizing things to my eyes when I tried to focus on it.<br> <br>When I heard a long sigh go up simultaneously from the watchers I risked a look again.<br> <br>There were two spheres floating where one had floated before.</p> </blockquote> Answer 18: <p>I think it is interesting to re-phrase this question relative to <strong>other</strong> axioms and/or theorems of ZFC. What Andrej Bauer's answer suggests is that it may not be the axiom of choice <em>per se</em> that is the culprit, but rather the underlying structure.</p> <p>For example, it is provable that the existence of non-Lebesgue measurable sets and the Banach–Tarski paradox are <strong>both</strong> an implication over ZF (without Choice) of the Hahn–Banach theorem (HB) in functional analysis. This means that we can prove those "pathologies" as a theorems from ZF+HB. </p> <p>Another way to put it is to analyze the structure within the context of different orders of logic. In ordinary first-order logic Choice is provably equivalent to the well-ordering theorem (WO) over ZF. What is different in second-order logic is that WO is strictly stronger than choice: WO $\vdash_{ZF}$ Choice, but Choice $\nvdash_{ZF}$ WO.</p> <p>In other words, one may get the feeling that perhaps there has been too much emphasis put on Choice. Other axioms and theorems can prove just as problematic. The approach described above is the essence of Reverse Mathematics (RM). While RM has been traditionally carried out at a much lower proof-theoretic strength level (subsystems of second-order arithmetic), in my opinion this provides a very useful framework for analyzing the foundations of other parts of mathematics.</p> Answer 19: <p>There are <em>two</em> ingredients in the Banach-Tarski decomposition theorem:</p> <ol> <li>The notion of <em>space</em>, together with derived notions of <em>part</em> and <em>decomposition</em>.</li> <li>The axiom of choice.</li> </ol> <p>Most discussion about the theorem revolve around the axiom of choice. I would like to point out that the notion of space can be put under scrutiny as well.</p> <p>The Banach-Tarski decomposition of the sphere produces non-measurable parts of the sphere. If we <em>restrict</em> the notion of &quot;part&quot; to &quot;measurable subset&quot; the theorem disappears. For instance, if we move over into a model of set theory (without choice) in which all sets are measurable, we will have no Banach-Tarski. This is all well known.</p> <p>Somewhat amazingly, we can make the Banach-Tarski decomposition go away by <em>extending</em> the notion of subspace, and keep choice too. Alex Simpson in <a href="https://web.archive.org/web/20160909042100/http://homepages.inf.ed.ac.uk/als/Research/Sources/mrs.pdf" rel="noreferrer">Measure, Randomness and Sublocales</a> (Annals of Pure and Applied Logic, 163(11), pp. 1642-1659, 2012) shows that this is achieved by generalizing the notion of topological space to that of <a href="https://ncatlab.org/nlab/show/locale" rel="noreferrer">locale</a>. He explains it thus:</p> <blockquote> <p>&quot;The different pieces in the partitions defined by Vitali and by Banach and Tarski are deeply intertangled with each other. According to our notion of “part”, two such intertangled pieces are not disjoint from each other, so additivity does not apply. An intuitive explanation for the failure of disjointness is that, although two such pieces share no point in common, they nevertheless overlap on the topological “glue” that bonds neighbouring points in <span class="math-container">$\mathbb{R}^n$</span> together.&quot;</p> </blockquote> <p>Peter Johnstone explained in <a href="http://pointlesstopology.com/the-point-of-pointless-topology.pdf" rel="noreferrer">The point of pointless topology</a> why locales have mathematical significance that goes far beyond fixing a strange theorem about decomposition of the sphere. Why isn't everyone using locales? I do not know, I think it is purely a historic accident. At some point in the 20th century mathematicians collectively lost the idea that there is more to space than just its points.</p> <p>I personally prefer to blame the trouble on the notion of space, rather than the axiom of choice. As far as possible, geometric problems should be the business of geometry, not logic or set theory. Mathematicians are used to operating with various kinds of spaces (in geometry, in analysis, in topology, in algebraic geometry, in computation, etc.) and so it seems only natural that one should worry about using the correct notion of space first, and about underlying foundational principles later. Good math is immune to changes in foundations.</p> Answer 20: <p>It's notable that most of the "bread and butter" mathematical consequences of the axiom of choice are actually consequences of <em>countable</em> choice. (Every infinite set contains a countable subset, a countable union of countable sets is countable, etc.) The Hahn-Banach theorem is a counterexample, but only if you want it for nonseparable spaces, and I can't think of any time I've ever needed this. When restricted to separable Banach spaces it doesn't require any choice principle at all! Whereas the seemingly pathological consequences of choice (existence of nonmeasurable sets, Banach-Tarski, well-ordering of the real line) generally do not follow from countable choice.</p> <p>So the argument from mathematical value seems to me to support countable choice more than full choice. But that isn't a very strong argument, is it? We can't decide whether an axiom is true based on whether we like its consequences. At best it's suggestive.</p> <p>Incidentally, I had the impression when I read Zermelo that he had great polemical skill, but none of his arguments seemed to get directly to the <em>truth</em> of the axiom. He argues for the mathematical value of the axiom. He points out that his critics have themselves on occasion unwittingly used the axiom, which is a devastating point, but has little bearing on the question of truth. (If I'm not mistaken, those unwitting uses were all of countable choice, by the way.)</p> <p>You ask if one should "believe" the axiom of choice, and I think you are right to put the word "believe" in quotes. I feel strongly that set-theoretic assertions are objectively meaningful, but I also feel that philosophers of mathematics have done a very poor job of clarifying what sets are. (Halmos: "A pack of wolves, a bunch of grapes, or a flock of pigeons are all examples of sets of things." Black: "It ought then to make sense, at least sometimes, to speak of being pursued by a set, or eating a set, or putting a set to flight.") If we can't even get that straight, it's hard to come to grips with questions about the truth of questionable axioms.</p> Answer 21: <p>The other answers don't seem to have said much about why the axiom of choice is widely regarded as plausible. Let me try to address that question.</p> <p>First let's dispose of some non-reasons. In response to your questions, I don't know of anyone who thinks that the Banach&ndash;Tarski paradox is a reason to believe in the axiom of choice. I also don't know of anyone who argues, "It is a priori plausible that there exist non-measurable sets; so the fact that the axiom of choice yields the attractive conclusion that there are non-measurable sets is a point in favor of believing the axiom of choice." Instead, those who are comfortable with the existence of non-measurable sets typically start by accepting the axiom of choice, and then they accept non-measurable sets as "part of the territory" that comes with the axiom of choice.</p> <p>Those who think that Banach&ndash;Tarski casts doubt on the axiom of choice typically have a philosophical predisposition that <i>math is supposed to model the physical world closely</i>. So for example, $\mathbb R^3$ is not just a random mathematical structure that we study purely for its own sake; it is supposed to be a decent model of physical space (or at least, open subsets of $\mathbb R^3$ are supposed to model localized regions of physical space). Banach&ndash;Tarski, when given a direct physical interpretation in this way, yields something that we "know" makes no physical sense, and so if we think that math is supposed to yield physical truth in this way, then Banach&ndash;Tarski is going to lead us to reject something in the math. Whether that "something" we reject is the axiom of choice is a separate question, and Andrej Bauer's excellent answer shows that there are other options, but the point I want to highlight is that we're going to be led down this path in the first place only if we have certain presuppositions about how math and physics are supposed to relate.</p> <p>There are others who don't view set theory in this way. According to them, set theory is supposed to be about <i>abstract collections of things</i>, and the way to arrive at axioms is by abstractly thinking about what properties they should have, not by comparing them with the physical world. The axiom of choice can be thought of as saying that if you have a bunch of nonempty collections of things, then there is another collection of things that contains one element from each of your original nonempty collections. Stated this way, the principle sounds intuitively plausible, and I would argue that this intuitive plausibility is, at least implicitly, the main argument in the minds of most people who accept the axiom of choice. If this is the way you think, then non-measurable sets and Banach&ndash;Tarski are not going to dissuade you from accepting the axiom of choice. Those phenomena will just lead you to say that we can't arrive at physical predictions from mathematics in such a naive manner; instead, to do physics, we have to formulate <i>physical theories</i>. Math can of course help a lot with the construction of physical theories, but it's not as simple as just saying that the mathematical theory of $\mathbb R^3$ <i>is</i> our theory of physical space.</p> <p>These two options aren't the only options. The <a href="https://en.wikipedia.org/wiki/Solovay_model">work of Solovay</a> shows that you can, to a large extent, <a href="https://books.google.com/ngrams/graph?content=have+your+cake+and+eat%2Ceat+your+cake+and+have&amp;year_start=1840&amp;year_end=2000&amp;corpus=15&amp;smoothing=3&amp;share=&amp;direct_url=t1%3B%2Chave%20your%20cake%20and%20eat%3B%2Cc0%3B.t1%3B%2Ceat%20your%20cake%20and%20have%3B%2Cc0">have your cake and eat it too</a>, by working in a set-theoretic universe where all subsets of $\mathbb R^n$ are Lebesgue-measurable and a weakened, but still quite strong, version of the axiom of choice known as "dependent choice" is available. Why Solovay's model hasn't become more popular is not completely clear, but perhaps part of the reason is that it feels like a "compromise position," and the people in the two different camps above have not seen any need to migrate to that kind of compromise.</p> Answer 22: <blockquote> <p>Are there reasons why it is plausible (for physicists, philosophers, mathematicians) to believe that not all sets should be measurable?</p> </blockquote> <p>Yes. If every set of reals is Lebesgue measurable, then you can partition $\mathbb{R}$ into <em>more than continuum many</em> pairwise disjoint non-empty pieces. (See <a href="https://mathoverflow.net/a/22935/33039">this answer and comments</a> for the details.)</p> <p>Surely the Banach-Tarski paradox seems unintuitive. But having a set that can be broken up into more pieces than there originally were... is just wrong.</p> Answer 23: <p>Arguments from physics may not help. Here is Bryce DeWitt <a href="http://science.sciencemag.org/content/182/4113/705">reviewing</a> Stephen Hawking and G.F.R. Ellis using the axiom of choice in 1973:</p> <blockquote> <p>The book also contains one failure to distinguish between mathematics and physics that is actually serious. This is in the proof of the main theorem of chapter 7, that given a set of Cauchy data on a smooth spacelike hypersurface there exists a unique maximal development therefrom of Einstein’s empty-space equations. The proof, essentially due to Choquet-Bruhat and Geroch, makes use of the axiom of choice, in the guise of Zorn’s lemma. Now mathematicians may use this axiom if they wish, but it has no place in physics. Physicists are already stretching things, from an operational standpoint, in using the axiom of infinity.</p> <p>It is not a question here of resurrecting an old and out-of-date mathematical controversy. The simple fact is that the axiom of choice never is really needed except when dealing with sets and relations in non-constructive ways. Many remarkable and beautiful theorems can be proved only with its aid. But its irrelevance to physics should be evident from the fact that its denial, as Paul Cohen has shown us, is equally consistent with the other axioms of set theory. And these other axioms suffice for the constructions of the real numbers, Hilbert spaces, C* algebras, and pseudo-Riemannian manifolds–that is, of all the paraphernalia of theoretical physics.</p> <p>In “proving” the global Cauchy development theorem with the aid of Zorn’s lemma what one is actually doing is assuming that a “choice function” exists for every set of developments extending a given Cauchy development. This, of course, is begging the question. The physicist’s job is not done until he can show, by an explicit algorithm or construction, how one could in principle always select a member from every such set of developments. Failing that he has proved nothing.</p> </blockquote> <p>Some physicists want to use the axiom of choice, but some physicists don't.</p> Answer 24: <p>Just a quick addendum: the result was actually found by Felix Hausdorff and then repackaged in a more spectacular form by Banach and Tarski. Hausdorff' point was precisely to show that the axiom of choice leads to such unreasonable consequences that it should probably be avoided. This was a hot topic of discussion during the following decades, and it seems that the answer at large from the mathematical community is: ok, there are inconveniences, but the advantages of using the AC are superior and we prefer to have it available.</p> Answer 25: <p>I think one of the strongest arguments for the axiom of choice is that every model of ZF contains as an inner model a constructible universe $L$, and $AC$ is a theorem of the constructible universe. We have</p> <p>$$ ZF+(V=L) \vdash AC $$</p> <p>In other words, a necessary condition for asserting $\neg AC$ is to first assert "there exists a set that cannot be constructed" &mdash; that is, it requires positing the existence of additional structure above and beyond what is guaranteed by ZF.</p> <p>Thus, it seems clear to me that ZFC is by far the better choice for <em>foundations</em>. One may still wish to work with another set theoretic universe, but that's most appropriately done as the study of <em>additional structure</em> built atop of the foundations, not by rewriting the foundations themselves.</p> Answer 26: <p>Physical applications of the Banach–Tarski theorem were explored by <a href="https://archive.org/details/Startling_Stories_v18n03_1949-01" rel="noreferrer">Henry Kuttner, The Time Axis, <em>Startling Stories</em> 18:3 (January 1949), 13–82</a>. Some excerpts from <a href="https://archive.org/stream/Startling_Stories_v18n03_1949-01#page/n65/mode/2up" rel="noreferrer">pp. 66–67</a>:</p> <blockquote> <p>"'Professor Raphael M. Robinson of the University of California now shows that it is possible to divide a solid sphere into a minimum of five pieces and reassemble them to form two spheres of the same size as the original one. Two of the pieces are used to form one of the new spheres and three to form the other.<br> <br>"'Some of the pieces must necessarily be of such complicated structure that it is impossible to assign volume to them. Otherwise the sum of the volumes of the five pieces would have to be equal both to the volume of the original sphere and to the sum of the volumes of the two new spheres, which is twice as great.'"<br> <br>[. . . .]<br> <br>"This is it," he said.<br> <br>Even the crowd around the neural-web table thinned as the workers in the laboratory flocked around him to watch.<br> <br>He had a sphere about the size of a grapefruit, floating in mid-air above his table. He did things to it with quick flashes of light that acted exactly like knives, in that it fell apart wherever the lights touched, but I got the impression that those divisions were much less simple than knife-cuts would be. The light shivered as it slashed and the cuts must have been very complex, dividing molecules with a selective precision beyond my powers of comprehension.<br> <br>The sphere floated apart. It changed shape under the lights. I am pretty sure it changed shape in four dimensions, because after a while I literally could not watch any more. The shape did agonizing things to my eyes when I tried to focus on it.<br> <br>When I heard a long sigh go up simultaneously from the watchers I risked a look again.<br> <br>There were two spheres floating where one had floated before.</p> </blockquote> Answer 27: <p>I think it is interesting to re-phrase this question relative to <strong>other</strong> axioms and/or theorems of ZFC. What Andrej Bauer's answer suggests is that it may not be the axiom of choice <em>per se</em> that is the culprit, but rather the underlying structure.</p> <p>For example, it is provable that the existence of non-Lebesgue measurable sets and the Banach–Tarski paradox are <strong>both</strong> an implication over ZF (without Choice) of the Hahn–Banach theorem (HB) in functional analysis. This means that we can prove those "pathologies" as a theorems from ZF+HB. </p> <p>Another way to put it is to analyze the structure within the context of different orders of logic. In ordinary first-order logic Choice is provably equivalent to the well-ordering theorem (WO) over ZF. What is different in second-order logic is that WO is strictly stronger than choice: WO $\vdash_{ZF}$ Choice, but Choice $\nvdash_{ZF}$ WO.</p> <p>In other words, one may get the feeling that perhaps there has been too much emphasis put on Choice. Other axioms and theorems can prove just as problematic. The approach described above is the essence of Reverse Mathematics (RM). While RM has been traditionally carried out at a much lower proof-theoretic strength level (subsystems of second-order arithmetic), in my opinion this provides a very useful framework for analyzing the foundations of other parts of mathematics.</p> Answer 28: <p>There are <em>two</em> ingredients in the Banach-Tarski decomposition theorem:</p> <ol> <li>The notion of <em>space</em>, together with derived notions of <em>part</em> and <em>decomposition</em>.</li> <li>The axiom of choice.</li> </ol> <p>Most discussion about the theorem revolve around the axiom of choice. I would like to point out that the notion of space can be put under scrutiny as well.</p> <p>The Banach-Tarski decomposition of the sphere produces non-measurable parts of the sphere. If we <em>restrict</em> the notion of &quot;part&quot; to &quot;measurable subset&quot; the theorem disappears. For instance, if we move over into a model of set theory (without choice) in which all sets are measurable, we will have no Banach-Tarski. This is all well known.</p> <p>Somewhat amazingly, we can make the Banach-Tarski decomposition go away by <em>extending</em> the notion of subspace, and keep choice too. Alex Simpson in <a href="https://web.archive.org/web/20160909042100/http://homepages.inf.ed.ac.uk/als/Research/Sources/mrs.pdf" rel="noreferrer">Measure, Randomness and Sublocales</a> (Annals of Pure and Applied Logic, 163(11), pp. 1642-1659, 2012) shows that this is achieved by generalizing the notion of topological space to that of <a href="https://ncatlab.org/nlab/show/locale" rel="noreferrer">locale</a>. He explains it thus:</p> <blockquote> <p>&quot;The different pieces in the partitions defined by Vitali and by Banach and Tarski are deeply intertangled with each other. According to our notion of “part”, two such intertangled pieces are not disjoint from each other, so additivity does not apply. An intuitive explanation for the failure of disjointness is that, although two such pieces share no point in common, they nevertheless overlap on the topological “glue” that bonds neighbouring points in <span class="math-container">$\mathbb{R}^n$</span> together.&quot;</p> </blockquote> <p>Peter Johnstone explained in <a href="http://pointlesstopology.com/the-point-of-pointless-topology.pdf" rel="noreferrer">The point of pointless topology</a> why locales have mathematical significance that goes far beyond fixing a strange theorem about decomposition of the sphere. Why isn't everyone using locales? I do not know, I think it is purely a historic accident. At some point in the 20th century mathematicians collectively lost the idea that there is more to space than just its points.</p> <p>I personally prefer to blame the trouble on the notion of space, rather than the axiom of choice. As far as possible, geometric problems should be the business of geometry, not logic or set theory. Mathematicians are used to operating with various kinds of spaces (in geometry, in analysis, in topology, in algebraic geometry, in computation, etc.) and so it seems only natural that one should worry about using the correct notion of space first, and about underlying foundational principles later. Good math is immune to changes in foundations.</p> Answer 29: <p>It's notable that most of the "bread and butter" mathematical consequences of the axiom of choice are actually consequences of <em>countable</em> choice. (Every infinite set contains a countable subset, a countable union of countable sets is countable, etc.) The Hahn-Banach theorem is a counterexample, but only if you want it for nonseparable spaces, and I can't think of any time I've ever needed this. When restricted to separable Banach spaces it doesn't require any choice principle at all! Whereas the seemingly pathological consequences of choice (existence of nonmeasurable sets, Banach-Tarski, well-ordering of the real line) generally do not follow from countable choice.</p> <p>So the argument from mathematical value seems to me to support countable choice more than full choice. But that isn't a very strong argument, is it? We can't decide whether an axiom is true based on whether we like its consequences. At best it's suggestive.</p> <p>Incidentally, I had the impression when I read Zermelo that he had great polemical skill, but none of his arguments seemed to get directly to the <em>truth</em> of the axiom. He argues for the mathematical value of the axiom. He points out that his critics have themselves on occasion unwittingly used the axiom, which is a devastating point, but has little bearing on the question of truth. (If I'm not mistaken, those unwitting uses were all of countable choice, by the way.)</p> <p>You ask if one should "believe" the axiom of choice, and I think you are right to put the word "believe" in quotes. I feel strongly that set-theoretic assertions are objectively meaningful, but I also feel that philosophers of mathematics have done a very poor job of clarifying what sets are. (Halmos: "A pack of wolves, a bunch of grapes, or a flock of pigeons are all examples of sets of things." Black: "It ought then to make sense, at least sometimes, to speak of being pursued by a set, or eating a set, or putting a set to flight.") If we can't even get that straight, it's hard to come to grips with questions about the truth of questionable axioms.</p> Answer 30: <p>The other answers don't seem to have said much about why the axiom of choice is widely regarded as plausible. Let me try to address that question.</p> <p>First let's dispose of some non-reasons. In response to your questions, I don't know of anyone who thinks that the Banach&ndash;Tarski paradox is a reason to believe in the axiom of choice. I also don't know of anyone who argues, "It is a priori plausible that there exist non-measurable sets; so the fact that the axiom of choice yields the attractive conclusion that there are non-measurable sets is a point in favor of believing the axiom of choice." Instead, those who are comfortable with the existence of non-measurable sets typically start by accepting the axiom of choice, and then they accept non-measurable sets as "part of the territory" that comes with the axiom of choice.</p> <p>Those who think that Banach&ndash;Tarski casts doubt on the axiom of choice typically have a philosophical predisposition that <i>math is supposed to model the physical world closely</i>. So for example, $\mathbb R^3$ is not just a random mathematical structure that we study purely for its own sake; it is supposed to be a decent model of physical space (or at least, open subsets of $\mathbb R^3$ are supposed to model localized regions of physical space). Banach&ndash;Tarski, when given a direct physical interpretation in this way, yields something that we "know" makes no physical sense, and so if we think that math is supposed to yield physical truth in this way, then Banach&ndash;Tarski is going to lead us to reject something in the math. Whether that "something" we reject is the axiom of choice is a separate question, and Andrej Bauer's excellent answer shows that there are other options, but the point I want to highlight is that we're going to be led down this path in the first place only if we have certain presuppositions about how math and physics are supposed to relate.</p> <p>There are others who don't view set theory in this way. According to them, set theory is supposed to be about <i>abstract collections of things</i>, and the way to arrive at axioms is by abstractly thinking about what properties they should have, not by comparing them with the physical world. The axiom of choice can be thought of as saying that if you have a bunch of nonempty collections of things, then there is another collection of things that contains one element from each of your original nonempty collections. Stated this way, the principle sounds intuitively plausible, and I would argue that this intuitive plausibility is, at least implicitly, the main argument in the minds of most people who accept the axiom of choice. If this is the way you think, then non-measurable sets and Banach&ndash;Tarski are not going to dissuade you from accepting the axiom of choice. Those phenomena will just lead you to say that we can't arrive at physical predictions from mathematics in such a naive manner; instead, to do physics, we have to formulate <i>physical theories</i>. Math can of course help a lot with the construction of physical theories, but it's not as simple as just saying that the mathematical theory of $\mathbb R^3$ <i>is</i> our theory of physical space.</p> <p>These two options aren't the only options. The <a href="https://en.wikipedia.org/wiki/Solovay_model">work of Solovay</a> shows that you can, to a large extent, <a href="https://books.google.com/ngrams/graph?content=have+your+cake+and+eat%2Ceat+your+cake+and+have&amp;year_start=1840&amp;year_end=2000&amp;corpus=15&amp;smoothing=3&amp;share=&amp;direct_url=t1%3B%2Chave%20your%20cake%20and%20eat%3B%2Cc0%3B.t1%3B%2Ceat%20your%20cake%20and%20have%3B%2Cc0">have your cake and eat it too</a>, by working in a set-theoretic universe where all subsets of $\mathbb R^n$ are Lebesgue-measurable and a weakened, but still quite strong, version of the axiom of choice known as "dependent choice" is available. Why Solovay's model hasn't become more popular is not completely clear, but perhaps part of the reason is that it feels like a "compromise position," and the people in the two different camps above have not seen any need to migrate to that kind of compromise.</p> Answer 31: <blockquote> <p>Are there reasons why it is plausible (for physicists, philosophers, mathematicians) to believe that not all sets should be measurable?</p> </blockquote> <p>Yes. If every set of reals is Lebesgue measurable, then you can partition $\mathbb{R}$ into <em>more than continuum many</em> pairwise disjoint non-empty pieces. (See <a href="https://mathoverflow.net/a/22935/33039">this answer and comments</a> for the details.)</p> <p>Surely the Banach-Tarski paradox seems unintuitive. But having a set that can be broken up into more pieces than there originally were... is just wrong.</p> Answer 32: <p>Arguments from physics may not help. Here is Bryce DeWitt <a href="http://science.sciencemag.org/content/182/4113/705">reviewing</a> Stephen Hawking and G.F.R. Ellis using the axiom of choice in 1973:</p> <blockquote> <p>The book also contains one failure to distinguish between mathematics and physics that is actually serious. This is in the proof of the main theorem of chapter 7, that given a set of Cauchy data on a smooth spacelike hypersurface there exists a unique maximal development therefrom of Einstein’s empty-space equations. The proof, essentially due to Choquet-Bruhat and Geroch, makes use of the axiom of choice, in the guise of Zorn’s lemma. Now mathematicians may use this axiom if they wish, but it has no place in physics. Physicists are already stretching things, from an operational standpoint, in using the axiom of infinity.</p> <p>It is not a question here of resurrecting an old and out-of-date mathematical controversy. The simple fact is that the axiom of choice never is really needed except when dealing with sets and relations in non-constructive ways. Many remarkable and beautiful theorems can be proved only with its aid. But its irrelevance to physics should be evident from the fact that its denial, as Paul Cohen has shown us, is equally consistent with the other axioms of set theory. And these other axioms suffice for the constructions of the real numbers, Hilbert spaces, C* algebras, and pseudo-Riemannian manifolds–that is, of all the paraphernalia of theoretical physics.</p> <p>In “proving” the global Cauchy development theorem with the aid of Zorn’s lemma what one is actually doing is assuming that a “choice function” exists for every set of developments extending a given Cauchy development. This, of course, is begging the question. The physicist’s job is not done until he can show, by an explicit algorithm or construction, how one could in principle always select a member from every such set of developments. Failing that he has proved nothing.</p> </blockquote> <p>Some physicists want to use the axiom of choice, but some physicists don't.</p> Answer 33: <p>Just a quick addendum: the result was actually found by Felix Hausdorff and then repackaged in a more spectacular form by Banach and Tarski. Hausdorff' point was precisely to show that the axiom of choice leads to such unreasonable consequences that it should probably be avoided. This was a hot topic of discussion during the following decades, and it seems that the answer at large from the mathematical community is: ok, there are inconveniences, but the advantages of using the AC are superior and we prefer to have it available.</p> Answer 34: <p>I think one of the strongest arguments for the axiom of choice is that every model of ZF contains as an inner model a constructible universe $L$, and $AC$ is a theorem of the constructible universe. We have</p> <p>$$ ZF+(V=L) \vdash AC $$</p> <p>In other words, a necessary condition for asserting $\neg AC$ is to first assert "there exists a set that cannot be constructed" &mdash; that is, it requires positing the existence of additional structure above and beyond what is guaranteed by ZF.</p> <p>Thus, it seems clear to me that ZFC is by far the better choice for <em>foundations</em>. One may still wish to work with another set theoretic universe, but that's most appropriately done as the study of <em>additional structure</em> built atop of the foundations, not by rewriting the foundations themselves.</p> Answer 35: <p>Physical applications of the Banach–Tarski theorem were explored by <a href="https://archive.org/details/Startling_Stories_v18n03_1949-01" rel="noreferrer">Henry Kuttner, The Time Axis, <em>Startling Stories</em> 18:3 (January 1949), 13–82</a>. Some excerpts from <a href="https://archive.org/stream/Startling_Stories_v18n03_1949-01#page/n65/mode/2up" rel="noreferrer">pp. 66–67</a>:</p> <blockquote> <p>"'Professor Raphael M. Robinson of the University of California now shows that it is possible to divide a solid sphere into a minimum of five pieces and reassemble them to form two spheres of the same size as the original one. Two of the pieces are used to form one of the new spheres and three to form the other.<br> <br>"'Some of the pieces must necessarily be of such complicated structure that it is impossible to assign volume to them. Otherwise the sum of the volumes of the five pieces would have to be equal both to the volume of the original sphere and to the sum of the volumes of the two new spheres, which is twice as great.'"<br> <br>[. . . .]<br> <br>"This is it," he said.<br> <br>Even the crowd around the neural-web table thinned as the workers in the laboratory flocked around him to watch.<br> <br>He had a sphere about the size of a grapefruit, floating in mid-air above his table. He did things to it with quick flashes of light that acted exactly like knives, in that it fell apart wherever the lights touched, but I got the impression that those divisions were much less simple than knife-cuts would be. The light shivered as it slashed and the cuts must have been very complex, dividing molecules with a selective precision beyond my powers of comprehension.<br> <br>The sphere floated apart. It changed shape under the lights. I am pretty sure it changed shape in four dimensions, because after a while I literally could not watch any more. The shape did agonizing things to my eyes when I tried to focus on it.<br> <br>When I heard a long sigh go up simultaneously from the watchers I risked a look again.<br> <br>There were two spheres floating where one had floated before.</p> </blockquote> Answer 36: <p>I think it is interesting to re-phrase this question relative to <strong>other</strong> axioms and/or theorems of ZFC. What Andrej Bauer's answer suggests is that it may not be the axiom of choice <em>per se</em> that is the culprit, but rather the underlying structure.</p> <p>For example, it is provable that the existence of non-Lebesgue measurable sets and the Banach–Tarski paradox are <strong>both</strong> an implication over ZF (without Choice) of the Hahn–Banach theorem (HB) in functional analysis. This means that we can prove those "pathologies" as a theorems from ZF+HB. </p> <p>Another way to put it is to analyze the structure within the context of different orders of logic. In ordinary first-order logic Choice is provably equivalent to the well-ordering theorem (WO) over ZF. What is different in second-order logic is that WO is strictly stronger than choice: WO $\vdash_{ZF}$ Choice, but Choice $\nvdash_{ZF}$ WO.</p> <p>In other words, one may get the feeling that perhaps there has been too much emphasis put on Choice. Other axioms and theorems can prove just as problematic. The approach described above is the essence of Reverse Mathematics (RM). While RM has been traditionally carried out at a much lower proof-theoretic strength level (subsystems of second-order arithmetic), in my opinion this provides a very useful framework for analyzing the foundations of other parts of mathematics.</p>
What is precisely still missing in Connes&#39; approach to RH?
259,263
54
https://mathoverflow.net/questions/259263/what-is-precisely-still-missing-in-connes-approach-to-rh
<p>I have read Connes' survey article <a href="http://www.alainconnes.org/docs/rhfinal.pdf" rel="noreferrer">http://www.alainconnes.org/docs/rhfinal.pdf</a> and I am somewhat familiar with his classic paper on the trace formula: <a href="http://www.alainconnes.org/docs/selecta.ps" rel="noreferrer">http://www.alainconnes.org/docs/selecta.ps</a></p> <p>Very roughly speaking the idea is to describe a dictionary which translates the concepts and techniques used in the proof of the analogue of the Riemann Hypothesis for function fields. This translation uses various techniques from tropical geometry and topos theory. At first I was hopeful I might understand the key issues with this translation, since I have some experience with the theory of Grothendieck toposes (or topoi). Nevertheless I have been lost when it comes to understanding precisely what the remaining problems are. As already briefly discussed in this thread: <a href="https://mathoverflow.net/questions/69389/riemann-hypothesis-via-absolute-geometry">Riemann hypothesis via absolute geometry</a> in the proof of the Riemann hypothesis for a curve $X$ over $F_p$ the surface $X \times X$ plays an important role. According to a new preprint of Connes / Consani there is now a suitable analogy for the surface $X \times X$ which is given by a topos called the "scaling site", cf. <a href="https://arxiv.org/abs/1603.03191" rel="noreferrer">https://arxiv.org/abs/1603.03191</a> </p> <p><strong>I would like to know what are the issues that are left open to complete the analogy with the proof of RH in the case of a curve over $F_p$?</strong></p>
[ { "answer_id": 259301, "score": 70, "text": "<p>First, recall the step's of Weil's proof, other than defining the surface:</p>\n\n<ol>\n<li>Develop an intersection theory of curves on surfaces.</li>\n<li>Show that the intersection of two specially chosen curves is equal to a coefficient of the zeta function.</li>\n<li>Prove the Hodge index theorem using the Riemann-Roch theorem for surfaces (or is there another proof?).</li>\n<li>By playing around with the two curves from step 2 and a few other simple curves, deduce an inequality for each coefficient of the logarithmic derivative of the zeta function.</li>\n<li>Conclude the Riemann hypothesis for the roots of the zeta function.</li>\n</ol>\n\n<p>Any approach which closely follows Weil's strategy would have to find analogues of these 5 steps.</p>\n\n<p>I think most people would say the crux of the issue is step 3 - it would be very interesting if some abstract/categorical/algebraic method produced any inequality involving the Riemann zeta function or related functions that was not previously known, even if it was short of the full Riemann hypothesis. Equally, many people fear that if we generalize too far from the world of curves and surfaces that we know and love, we might lose the ability to prove concrete inequalities like this. nfdc23 discussed this issue in the comments.</p>\n\n<hr>\n\n<p>In the survey article you linke, Connes says:</p>\n\n<blockquote>\n <p>At this point, what is missing is an intersection theory and a Riemann-Roch theorem on the square of the arithmetic site.</p>\n</blockquote>\n\n<p>I think this is still the case - the Riemann-Roch theorem in the paper you link seems to be for certain curves in the scaling site. (If it were for the surface, I wouldn't expect to see it on the arXiv until the Riemann hypothesis were deduced from it, or a very convincing argument was found that it was not sufficient to prove the Riemann hypothesis!)</p>\n\n<hr>\n\n<p>I'm not sure how helpful knowledge of toposes (alone) will be for understanding this issue. The reason is that the key parts of any definition of intersection theory of curves on surfaces do not involve very much the topos structure of the surface, but rather other structures. </p>\n\n<p>I know of a few different perspectives on intersection theory:</p>\n\n<ol>\n<li><p>Add ample classes, apply a moving lemma, and then concretely count the intersections of curves. Here the main issues are global geometric considerations - ample divisors and their relationships to projective embeddings and hyperplanes and so on. I don't think toposes are a very useful tool for understanding these things.</p></li>\n<li><p>Apply local formulas for the intersection number, and then sum over the point of intersection. This is more promising as e.g. Serre's formula is cohomological in nature, but it's the wrong kind of cohomology - for modules, not for toposes. I think it might be hard to define the right cohomology groups over semirings.</p></li>\n<li><p>Take the cup product of their classes in a suitable cohomology theory. This at first looks promising, because one of our possible choices for a cohomology theory, etale cohomology, is defined using a topos. Unfortunately, because etale cohomology is naturally l-adic, it is difficult to establish positivity from it. Of course one could look to generalizing Deligne's proof of the Riemann hypothesis, but that is a different matter and much more complicated. (Peter Sarnak has suggested that Deligne's proof provides a guide to how one should try to prove the Riemann hypothesis for number fields, but it is not the etale topos but rather the use of fmailies in the proof that he wants to mimic.)</p></li>\n<li><p>View one curve as an honest curve and the other as the divisor of a line bundle and pull the line bundle back to the curve, then compute its degree. This is, I think, the easiest way to define the intersection number, but one needs some of the other characterizations to prove its properties (even symmetry). This looks to me like it is at least potentially meaningful in some topos and I don't know exactly to what extent it can be applied to the arithmetic site.</p></li>\n</ol>\n\n<p>Remember that even among locally ringed spaces, the ones where you have a good intersection theory are very specific (If you want to intersect arbitrary closed subsets I think you need smooth proper varieties over a field). For locally semiringed topoi, the situation is presumably much worse.</p>\n\n<p>No one is saying it is impossible to develop an applicable intersection theory, but I don't think anyone really knows how to begin -- if they did, I'm sure they would be working furiously on it.</p>\n" }, { "answer_id": 259388, "score": 24, "text": "<p>There is a very interesting book which appeared recently precisely on this topic: <a href=\"https://books.google.com/books?id=RbhEAgAAQBAJ&amp;printsec=frontcover&amp;source=gbs_ge_summary_r&amp;cad=0#v=onepage&amp;q&amp;f=false\" rel=\"noreferrer\">\"The Riemann Hypothesis for Function Fields\"</a> by Machiel van Frankenhuijsen. His attempt at answering the OP's question about what is missing is in the last two chapters.</p>\n\n<p>My impression is that Connes' work on RH, like that of great painters such as Picasso, can be divided into different periods. I would say the first period is\nabout exploring connections to statistical mechanics and the notion of symmetry breaking. The main result is his article with Bost on the now called Bost-Connes system. However, this did not yet contain an approach towards RH.\nThe second period is that of his Selecta paper with the trace formula.\nFinally the third period involves having \"fun with $\\mathbb{F}_1$\", hyperreals, toposes, etc. The third period is about constructing suitable objects which would allow one to mimic Weil's proof. I am not familiar with this last development, but I think Will did a good job at explaining what is missing regarding the approach from this third period. The book I mentioned and the little I know concern the second period.</p>\n\n<p>Let me first give a general somewhat vague opinion: if one has a reasonably clear description of what $A$ and $B$ are, then it is easier to prove a conjecture which says $A=B$ than one which says $A\\neq B$ (or $A&gt;B$, $A\\ge B$, etc.). For example, I think the <a href=\"https://arxiv.org/abs/1508.07205\" rel=\"noreferrer\">recent work</a> by Kronheimer and Mrowka is significant progress on the four-color theorem. The latter says $A&gt;B$ where\n$B=0$ and $A$ is the number of Tait colorings. They produced some invariant $C$\nwhich they showed satisfies $C&gt;B$ and reduced the problem to the conjecture $A=C$. Connes' Selecta article reduced Weil positivity (an $A\\ge B$ type problem) and thus RH to an $A=B$ type problem, i.e., proving his global trace formula which is formally similar to a theorem of Guillemin and Sternberg for flows on Riemannian manifolds. He also proved the semi-local analogue of his formula. Take a finite group $G$ with a unitary representation $U$ and consider the matrix ${\\rm tr}(U(gh^{-1}))$ with rows and columns indexed by $g,h\\in G$.\nThis matrix is positive semidefinite. Weil positivity follows from Connes' global trace formula for essentially the same reason. </p>\n\n<p>Proving the trace formula, prima facie, looks like a problem of harmonic analysis or even constructive quantum field theory since one has a formally precise but still physicsy looking infinite-dimensional trace to make sense of by introducing UV and IR cutoffs and then analyzing the limit of removing them. There has been follow-ups by Haran (<a href=\"https://books.google.com/books?hl=en&amp;lr=&amp;id=eGIbCAAAQBAJ&amp;oi=fnd&amp;pg=PA93&amp;ots=YkfLo5jT04&amp;sig=lqQNwqZUfh_obuH8YVvY4IeYvrg&amp;redir_esc=y#v=onepage&amp;q&amp;f=false\" rel=\"noreferrer\">here</a> and <a href=\"https://books.google.com/books?id=bsUkkEraBvAC&amp;printsec=frontcover#v=onepage&amp;q&amp;f=false\" rel=\"noreferrer\">here</a>) and Burnol\n(<a href=\"https://arxiv.org/abs/math/0101068\" rel=\"noreferrer\">here</a> and <a href=\"https://arxiv.org/abs/math/0112254\" rel=\"noreferrer\">here</a>) but, as far as I know, the trace formula has not been proved\nin this approach, even in the function field case where at least we know it should be true.</p>\n" }, { "answer_id": 277076, "score": 4, "text": "<p>Terence Tao answered this question as follows:</p>\n\n<p>\" He (Connes) also discussed his recent paper with Consani in which they managed to view the (completed) Riemann zeta function as the Hasse-Weil zeta function of a curve over F_1, where F_1 is interpreted as the tropical ring {0,1}. So the Riemann hypothesis for the integers is now expressed in a very similar fashion to the Riemann hypothesis for curves, which has a number of proofs. Unfortunately, they are missing a huge ingredient in the dictionary, namely they have no Riemann-Roch theorem over F_1. Still, it is a very suggestive similarity, and something to keep an eye on in the future... \"</p>\n" }, { "answer_id": 259301, "score": 70, "text": "<p>First, recall the step's of Weil's proof, other than defining the surface:</p>\n\n<ol>\n<li>Develop an intersection theory of curves on surfaces.</li>\n<li>Show that the intersection of two specially chosen curves is equal to a coefficient of the zeta function.</li>\n<li>Prove the Hodge index theorem using the Riemann-Roch theorem for surfaces (or is there another proof?).</li>\n<li>By playing around with the two curves from step 2 and a few other simple curves, deduce an inequality for each coefficient of the logarithmic derivative of the zeta function.</li>\n<li>Conclude the Riemann hypothesis for the roots of the zeta function.</li>\n</ol>\n\n<p>Any approach which closely follows Weil's strategy would have to find analogues of these 5 steps.</p>\n\n<p>I think most people would say the crux of the issue is step 3 - it would be very interesting if some abstract/categorical/algebraic method produced any inequality involving the Riemann zeta function or related functions that was not previously known, even if it was short of the full Riemann hypothesis. Equally, many people fear that if we generalize too far from the world of curves and surfaces that we know and love, we might lose the ability to prove concrete inequalities like this. nfdc23 discussed this issue in the comments.</p>\n\n<hr>\n\n<p>In the survey article you linke, Connes says:</p>\n\n<blockquote>\n <p>At this point, what is missing is an intersection theory and a Riemann-Roch theorem on the square of the arithmetic site.</p>\n</blockquote>\n\n<p>I think this is still the case - the Riemann-Roch theorem in the paper you link seems to be for certain curves in the scaling site. (If it were for the surface, I wouldn't expect to see it on the arXiv until the Riemann hypothesis were deduced from it, or a very convincing argument was found that it was not sufficient to prove the Riemann hypothesis!)</p>\n\n<hr>\n\n<p>I'm not sure how helpful knowledge of toposes (alone) will be for understanding this issue. The reason is that the key parts of any definition of intersection theory of curves on surfaces do not involve very much the topos structure of the surface, but rather other structures. </p>\n\n<p>I know of a few different perspectives on intersection theory:</p>\n\n<ol>\n<li><p>Add ample classes, apply a moving lemma, and then concretely count the intersections of curves. Here the main issues are global geometric considerations - ample divisors and their relationships to projective embeddings and hyperplanes and so on. I don't think toposes are a very useful tool for understanding these things.</p></li>\n<li><p>Apply local formulas for the intersection number, and then sum over the point of intersection. This is more promising as e.g. Serre's formula is cohomological in nature, but it's the wrong kind of cohomology - for modules, not for toposes. I think it might be hard to define the right cohomology groups over semirings.</p></li>\n<li><p>Take the cup product of their classes in a suitable cohomology theory. This at first looks promising, because one of our possible choices for a cohomology theory, etale cohomology, is defined using a topos. Unfortunately, because etale cohomology is naturally l-adic, it is difficult to establish positivity from it. Of course one could look to generalizing Deligne's proof of the Riemann hypothesis, but that is a different matter and much more complicated. (Peter Sarnak has suggested that Deligne's proof provides a guide to how one should try to prove the Riemann hypothesis for number fields, but it is not the etale topos but rather the use of fmailies in the proof that he wants to mimic.)</p></li>\n<li><p>View one curve as an honest curve and the other as the divisor of a line bundle and pull the line bundle back to the curve, then compute its degree. This is, I think, the easiest way to define the intersection number, but one needs some of the other characterizations to prove its properties (even symmetry). This looks to me like it is at least potentially meaningful in some topos and I don't know exactly to what extent it can be applied to the arithmetic site.</p></li>\n</ol>\n\n<p>Remember that even among locally ringed spaces, the ones where you have a good intersection theory are very specific (If you want to intersect arbitrary closed subsets I think you need smooth proper varieties over a field). For locally semiringed topoi, the situation is presumably much worse.</p>\n\n<p>No one is saying it is impossible to develop an applicable intersection theory, but I don't think anyone really knows how to begin -- if they did, I'm sure they would be working furiously on it.</p>\n" }, { "answer_id": 259388, "score": 24, "text": "<p>There is a very interesting book which appeared recently precisely on this topic: <a href=\"https://books.google.com/books?id=RbhEAgAAQBAJ&amp;printsec=frontcover&amp;source=gbs_ge_summary_r&amp;cad=0#v=onepage&amp;q&amp;f=false\" rel=\"noreferrer\">\"The Riemann Hypothesis for Function Fields\"</a> by Machiel van Frankenhuijsen. His attempt at answering the OP's question about what is missing is in the last two chapters.</p>\n\n<p>My impression is that Connes' work on RH, like that of great painters such as Picasso, can be divided into different periods. I would say the first period is\nabout exploring connections to statistical mechanics and the notion of symmetry breaking. The main result is his article with Bost on the now called Bost-Connes system. However, this did not yet contain an approach towards RH.\nThe second period is that of his Selecta paper with the trace formula.\nFinally the third period involves having \"fun with $\\mathbb{F}_1$\", hyperreals, toposes, etc. The third period is about constructing suitable objects which would allow one to mimic Weil's proof. I am not familiar with this last development, but I think Will did a good job at explaining what is missing regarding the approach from this third period. The book I mentioned and the little I know concern the second period.</p>\n\n<p>Let me first give a general somewhat vague opinion: if one has a reasonably clear description of what $A$ and $B$ are, then it is easier to prove a conjecture which says $A=B$ than one which says $A\\neq B$ (or $A&gt;B$, $A\\ge B$, etc.). For example, I think the <a href=\"https://arxiv.org/abs/1508.07205\" rel=\"noreferrer\">recent work</a> by Kronheimer and Mrowka is significant progress on the four-color theorem. The latter says $A&gt;B$ where\n$B=0$ and $A$ is the number of Tait colorings. They produced some invariant $C$\nwhich they showed satisfies $C&gt;B$ and reduced the problem to the conjecture $A=C$. Connes' Selecta article reduced Weil positivity (an $A\\ge B$ type problem) and thus RH to an $A=B$ type problem, i.e., proving his global trace formula which is formally similar to a theorem of Guillemin and Sternberg for flows on Riemannian manifolds. He also proved the semi-local analogue of his formula. Take a finite group $G$ with a unitary representation $U$ and consider the matrix ${\\rm tr}(U(gh^{-1}))$ with rows and columns indexed by $g,h\\in G$.\nThis matrix is positive semidefinite. Weil positivity follows from Connes' global trace formula for essentially the same reason. </p>\n\n<p>Proving the trace formula, prima facie, looks like a problem of harmonic analysis or even constructive quantum field theory since one has a formally precise but still physicsy looking infinite-dimensional trace to make sense of by introducing UV and IR cutoffs and then analyzing the limit of removing them. There has been follow-ups by Haran (<a href=\"https://books.google.com/books?hl=en&amp;lr=&amp;id=eGIbCAAAQBAJ&amp;oi=fnd&amp;pg=PA93&amp;ots=YkfLo5jT04&amp;sig=lqQNwqZUfh_obuH8YVvY4IeYvrg&amp;redir_esc=y#v=onepage&amp;q&amp;f=false\" rel=\"noreferrer\">here</a> and <a href=\"https://books.google.com/books?id=bsUkkEraBvAC&amp;printsec=frontcover#v=onepage&amp;q&amp;f=false\" rel=\"noreferrer\">here</a>) and Burnol\n(<a href=\"https://arxiv.org/abs/math/0101068\" rel=\"noreferrer\">here</a> and <a href=\"https://arxiv.org/abs/math/0112254\" rel=\"noreferrer\">here</a>) but, as far as I know, the trace formula has not been proved\nin this approach, even in the function field case where at least we know it should be true.</p>\n" }, { "answer_id": 277076, "score": 4, "text": "<p>Terence Tao answered this question as follows:</p>\n\n<p>\" He (Connes) also discussed his recent paper with Consani in which they managed to view the (completed) Riemann zeta function as the Hasse-Weil zeta function of a curve over F_1, where F_1 is interpreted as the tropical ring {0,1}. So the Riemann hypothesis for the integers is now expressed in a very similar fashion to the Riemann hypothesis for curves, which has a number of proofs. Unfortunately, they are missing a huge ingredient in the dictionary, namely they have no Riemann-Roch theorem over F_1. Still, it is a very suggestive similarity, and something to keep an eye on in the future... \"</p>\n" }, { "answer_id": 259301, "score": 70, "text": "<p>First, recall the step's of Weil's proof, other than defining the surface:</p>\n\n<ol>\n<li>Develop an intersection theory of curves on surfaces.</li>\n<li>Show that the intersection of two specially chosen curves is equal to a coefficient of the zeta function.</li>\n<li>Prove the Hodge index theorem using the Riemann-Roch theorem for surfaces (or is there another proof?).</li>\n<li>By playing around with the two curves from step 2 and a few other simple curves, deduce an inequality for each coefficient of the logarithmic derivative of the zeta function.</li>\n<li>Conclude the Riemann hypothesis for the roots of the zeta function.</li>\n</ol>\n\n<p>Any approach which closely follows Weil's strategy would have to find analogues of these 5 steps.</p>\n\n<p>I think most people would say the crux of the issue is step 3 - it would be very interesting if some abstract/categorical/algebraic method produced any inequality involving the Riemann zeta function or related functions that was not previously known, even if it was short of the full Riemann hypothesis. Equally, many people fear that if we generalize too far from the world of curves and surfaces that we know and love, we might lose the ability to prove concrete inequalities like this. nfdc23 discussed this issue in the comments.</p>\n\n<hr>\n\n<p>In the survey article you linke, Connes says:</p>\n\n<blockquote>\n <p>At this point, what is missing is an intersection theory and a Riemann-Roch theorem on the square of the arithmetic site.</p>\n</blockquote>\n\n<p>I think this is still the case - the Riemann-Roch theorem in the paper you link seems to be for certain curves in the scaling site. (If it were for the surface, I wouldn't expect to see it on the arXiv until the Riemann hypothesis were deduced from it, or a very convincing argument was found that it was not sufficient to prove the Riemann hypothesis!)</p>\n\n<hr>\n\n<p>I'm not sure how helpful knowledge of toposes (alone) will be for understanding this issue. The reason is that the key parts of any definition of intersection theory of curves on surfaces do not involve very much the topos structure of the surface, but rather other structures. </p>\n\n<p>I know of a few different perspectives on intersection theory:</p>\n\n<ol>\n<li><p>Add ample classes, apply a moving lemma, and then concretely count the intersections of curves. Here the main issues are global geometric considerations - ample divisors and their relationships to projective embeddings and hyperplanes and so on. I don't think toposes are a very useful tool for understanding these things.</p></li>\n<li><p>Apply local formulas for the intersection number, and then sum over the point of intersection. This is more promising as e.g. Serre's formula is cohomological in nature, but it's the wrong kind of cohomology - for modules, not for toposes. I think it might be hard to define the right cohomology groups over semirings.</p></li>\n<li><p>Take the cup product of their classes in a suitable cohomology theory. This at first looks promising, because one of our possible choices for a cohomology theory, etale cohomology, is defined using a topos. Unfortunately, because etale cohomology is naturally l-adic, it is difficult to establish positivity from it. Of course one could look to generalizing Deligne's proof of the Riemann hypothesis, but that is a different matter and much more complicated. (Peter Sarnak has suggested that Deligne's proof provides a guide to how one should try to prove the Riemann hypothesis for number fields, but it is not the etale topos but rather the use of fmailies in the proof that he wants to mimic.)</p></li>\n<li><p>View one curve as an honest curve and the other as the divisor of a line bundle and pull the line bundle back to the curve, then compute its degree. This is, I think, the easiest way to define the intersection number, but one needs some of the other characterizations to prove its properties (even symmetry). This looks to me like it is at least potentially meaningful in some topos and I don't know exactly to what extent it can be applied to the arithmetic site.</p></li>\n</ol>\n\n<p>Remember that even among locally ringed spaces, the ones where you have a good intersection theory are very specific (If you want to intersect arbitrary closed subsets I think you need smooth proper varieties over a field). For locally semiringed topoi, the situation is presumably much worse.</p>\n\n<p>No one is saying it is impossible to develop an applicable intersection theory, but I don't think anyone really knows how to begin -- if they did, I'm sure they would be working furiously on it.</p>\n" }, { "answer_id": 259388, "score": 24, "text": "<p>There is a very interesting book which appeared recently precisely on this topic: <a href=\"https://books.google.com/books?id=RbhEAgAAQBAJ&amp;printsec=frontcover&amp;source=gbs_ge_summary_r&amp;cad=0#v=onepage&amp;q&amp;f=false\" rel=\"noreferrer\">\"The Riemann Hypothesis for Function Fields\"</a> by Machiel van Frankenhuijsen. His attempt at answering the OP's question about what is missing is in the last two chapters.</p>\n\n<p>My impression is that Connes' work on RH, like that of great painters such as Picasso, can be divided into different periods. I would say the first period is\nabout exploring connections to statistical mechanics and the notion of symmetry breaking. The main result is his article with Bost on the now called Bost-Connes system. However, this did not yet contain an approach towards RH.\nThe second period is that of his Selecta paper with the trace formula.\nFinally the third period involves having \"fun with $\\mathbb{F}_1$\", hyperreals, toposes, etc. The third period is about constructing suitable objects which would allow one to mimic Weil's proof. I am not familiar with this last development, but I think Will did a good job at explaining what is missing regarding the approach from this third period. The book I mentioned and the little I know concern the second period.</p>\n\n<p>Let me first give a general somewhat vague opinion: if one has a reasonably clear description of what $A$ and $B$ are, then it is easier to prove a conjecture which says $A=B$ than one which says $A\\neq B$ (or $A&gt;B$, $A\\ge B$, etc.). For example, I think the <a href=\"https://arxiv.org/abs/1508.07205\" rel=\"noreferrer\">recent work</a> by Kronheimer and Mrowka is significant progress on the four-color theorem. The latter says $A&gt;B$ where\n$B=0$ and $A$ is the number of Tait colorings. They produced some invariant $C$\nwhich they showed satisfies $C&gt;B$ and reduced the problem to the conjecture $A=C$. Connes' Selecta article reduced Weil positivity (an $A\\ge B$ type problem) and thus RH to an $A=B$ type problem, i.e., proving his global trace formula which is formally similar to a theorem of Guillemin and Sternberg for flows on Riemannian manifolds. He also proved the semi-local analogue of his formula. Take a finite group $G$ with a unitary representation $U$ and consider the matrix ${\\rm tr}(U(gh^{-1}))$ with rows and columns indexed by $g,h\\in G$.\nThis matrix is positive semidefinite. Weil positivity follows from Connes' global trace formula for essentially the same reason. </p>\n\n<p>Proving the trace formula, prima facie, looks like a problem of harmonic analysis or even constructive quantum field theory since one has a formally precise but still physicsy looking infinite-dimensional trace to make sense of by introducing UV and IR cutoffs and then analyzing the limit of removing them. There has been follow-ups by Haran (<a href=\"https://books.google.com/books?hl=en&amp;lr=&amp;id=eGIbCAAAQBAJ&amp;oi=fnd&amp;pg=PA93&amp;ots=YkfLo5jT04&amp;sig=lqQNwqZUfh_obuH8YVvY4IeYvrg&amp;redir_esc=y#v=onepage&amp;q&amp;f=false\" rel=\"noreferrer\">here</a> and <a href=\"https://books.google.com/books?id=bsUkkEraBvAC&amp;printsec=frontcover#v=onepage&amp;q&amp;f=false\" rel=\"noreferrer\">here</a>) and Burnol\n(<a href=\"https://arxiv.org/abs/math/0101068\" rel=\"noreferrer\">here</a> and <a href=\"https://arxiv.org/abs/math/0112254\" rel=\"noreferrer\">here</a>) but, as far as I know, the trace formula has not been proved\nin this approach, even in the function field case where at least we know it should be true.</p>\n" }, { "answer_id": 277076, "score": 4, "text": "<p>Terence Tao answered this question as follows:</p>\n\n<p>\" He (Connes) also discussed his recent paper with Consani in which they managed to view the (completed) Riemann zeta function as the Hasse-Weil zeta function of a curve over F_1, where F_1 is interpreted as the tropical ring {0,1}. So the Riemann hypothesis for the integers is now expressed in a very similar fashion to the Riemann hypothesis for curves, which has a number of proofs. Unfortunately, they are missing a huge ingredient in the dictionary, namely they have no Riemann-Roch theorem over F_1. Still, it is a very suggestive similarity, and something to keep an eye on in the future... \"</p>\n" }, { "answer_id": 259301, "score": 70, "text": "<p>First, recall the step's of Weil's proof, other than defining the surface:</p>\n\n<ol>\n<li>Develop an intersection theory of curves on surfaces.</li>\n<li>Show that the intersection of two specially chosen curves is equal to a coefficient of the zeta function.</li>\n<li>Prove the Hodge index theorem using the Riemann-Roch theorem for surfaces (or is there another proof?).</li>\n<li>By playing around with the two curves from step 2 and a few other simple curves, deduce an inequality for each coefficient of the logarithmic derivative of the zeta function.</li>\n<li>Conclude the Riemann hypothesis for the roots of the zeta function.</li>\n</ol>\n\n<p>Any approach which closely follows Weil's strategy would have to find analogues of these 5 steps.</p>\n\n<p>I think most people would say the crux of the issue is step 3 - it would be very interesting if some abstract/categorical/algebraic method produced any inequality involving the Riemann zeta function or related functions that was not previously known, even if it was short of the full Riemann hypothesis. Equally, many people fear that if we generalize too far from the world of curves and surfaces that we know and love, we might lose the ability to prove concrete inequalities like this. nfdc23 discussed this issue in the comments.</p>\n\n<hr>\n\n<p>In the survey article you linke, Connes says:</p>\n\n<blockquote>\n <p>At this point, what is missing is an intersection theory and a Riemann-Roch theorem on the square of the arithmetic site.</p>\n</blockquote>\n\n<p>I think this is still the case - the Riemann-Roch theorem in the paper you link seems to be for certain curves in the scaling site. (If it were for the surface, I wouldn't expect to see it on the arXiv until the Riemann hypothesis were deduced from it, or a very convincing argument was found that it was not sufficient to prove the Riemann hypothesis!)</p>\n\n<hr>\n\n<p>I'm not sure how helpful knowledge of toposes (alone) will be for understanding this issue. The reason is that the key parts of any definition of intersection theory of curves on surfaces do not involve very much the topos structure of the surface, but rather other structures. </p>\n\n<p>I know of a few different perspectives on intersection theory:</p>\n\n<ol>\n<li><p>Add ample classes, apply a moving lemma, and then concretely count the intersections of curves. Here the main issues are global geometric considerations - ample divisors and their relationships to projective embeddings and hyperplanes and so on. I don't think toposes are a very useful tool for understanding these things.</p></li>\n<li><p>Apply local formulas for the intersection number, and then sum over the point of intersection. This is more promising as e.g. Serre's formula is cohomological in nature, but it's the wrong kind of cohomology - for modules, not for toposes. I think it might be hard to define the right cohomology groups over semirings.</p></li>\n<li><p>Take the cup product of their classes in a suitable cohomology theory. This at first looks promising, because one of our possible choices for a cohomology theory, etale cohomology, is defined using a topos. Unfortunately, because etale cohomology is naturally l-adic, it is difficult to establish positivity from it. Of course one could look to generalizing Deligne's proof of the Riemann hypothesis, but that is a different matter and much more complicated. (Peter Sarnak has suggested that Deligne's proof provides a guide to how one should try to prove the Riemann hypothesis for number fields, but it is not the etale topos but rather the use of fmailies in the proof that he wants to mimic.)</p></li>\n<li><p>View one curve as an honest curve and the other as the divisor of a line bundle and pull the line bundle back to the curve, then compute its degree. This is, I think, the easiest way to define the intersection number, but one needs some of the other characterizations to prove its properties (even symmetry). This looks to me like it is at least potentially meaningful in some topos and I don't know exactly to what extent it can be applied to the arithmetic site.</p></li>\n</ol>\n\n<p>Remember that even among locally ringed spaces, the ones where you have a good intersection theory are very specific (If you want to intersect arbitrary closed subsets I think you need smooth proper varieties over a field). For locally semiringed topoi, the situation is presumably much worse.</p>\n\n<p>No one is saying it is impossible to develop an applicable intersection theory, but I don't think anyone really knows how to begin -- if they did, I'm sure they would be working furiously on it.</p>\n" }, { "answer_id": 259388, "score": 24, "text": "<p>There is a very interesting book which appeared recently precisely on this topic: <a href=\"https://books.google.com/books?id=RbhEAgAAQBAJ&amp;printsec=frontcover&amp;source=gbs_ge_summary_r&amp;cad=0#v=onepage&amp;q&amp;f=false\" rel=\"noreferrer\">\"The Riemann Hypothesis for Function Fields\"</a> by Machiel van Frankenhuijsen. His attempt at answering the OP's question about what is missing is in the last two chapters.</p>\n\n<p>My impression is that Connes' work on RH, like that of great painters such as Picasso, can be divided into different periods. I would say the first period is\nabout exploring connections to statistical mechanics and the notion of symmetry breaking. The main result is his article with Bost on the now called Bost-Connes system. However, this did not yet contain an approach towards RH.\nThe second period is that of his Selecta paper with the trace formula.\nFinally the third period involves having \"fun with $\\mathbb{F}_1$\", hyperreals, toposes, etc. The third period is about constructing suitable objects which would allow one to mimic Weil's proof. I am not familiar with this last development, but I think Will did a good job at explaining what is missing regarding the approach from this third period. The book I mentioned and the little I know concern the second period.</p>\n\n<p>Let me first give a general somewhat vague opinion: if one has a reasonably clear description of what $A$ and $B$ are, then it is easier to prove a conjecture which says $A=B$ than one which says $A\\neq B$ (or $A&gt;B$, $A\\ge B$, etc.). For example, I think the <a href=\"https://arxiv.org/abs/1508.07205\" rel=\"noreferrer\">recent work</a> by Kronheimer and Mrowka is significant progress on the four-color theorem. The latter says $A&gt;B$ where\n$B=0$ and $A$ is the number of Tait colorings. They produced some invariant $C$\nwhich they showed satisfies $C&gt;B$ and reduced the problem to the conjecture $A=C$. Connes' Selecta article reduced Weil positivity (an $A\\ge B$ type problem) and thus RH to an $A=B$ type problem, i.e., proving his global trace formula which is formally similar to a theorem of Guillemin and Sternberg for flows on Riemannian manifolds. He also proved the semi-local analogue of his formula. Take a finite group $G$ with a unitary representation $U$ and consider the matrix ${\\rm tr}(U(gh^{-1}))$ with rows and columns indexed by $g,h\\in G$.\nThis matrix is positive semidefinite. Weil positivity follows from Connes' global trace formula for essentially the same reason. </p>\n\n<p>Proving the trace formula, prima facie, looks like a problem of harmonic analysis or even constructive quantum field theory since one has a formally precise but still physicsy looking infinite-dimensional trace to make sense of by introducing UV and IR cutoffs and then analyzing the limit of removing them. There has been follow-ups by Haran (<a href=\"https://books.google.com/books?hl=en&amp;lr=&amp;id=eGIbCAAAQBAJ&amp;oi=fnd&amp;pg=PA93&amp;ots=YkfLo5jT04&amp;sig=lqQNwqZUfh_obuH8YVvY4IeYvrg&amp;redir_esc=y#v=onepage&amp;q&amp;f=false\" rel=\"noreferrer\">here</a> and <a href=\"https://books.google.com/books?id=bsUkkEraBvAC&amp;printsec=frontcover#v=onepage&amp;q&amp;f=false\" rel=\"noreferrer\">here</a>) and Burnol\n(<a href=\"https://arxiv.org/abs/math/0101068\" rel=\"noreferrer\">here</a> and <a href=\"https://arxiv.org/abs/math/0112254\" rel=\"noreferrer\">here</a>) but, as far as I know, the trace formula has not been proved\nin this approach, even in the function field case where at least we know it should be true.</p>\n" }, { "answer_id": 277076, "score": 4, "text": "<p>Terence Tao answered this question as follows:</p>\n\n<p>\" He (Connes) also discussed his recent paper with Consani in which they managed to view the (completed) Riemann zeta function as the Hasse-Weil zeta function of a curve over F_1, where F_1 is interpreted as the tropical ring {0,1}. So the Riemann hypothesis for the integers is now expressed in a very similar fashion to the Riemann hypothesis for curves, which has a number of proofs. Unfortunately, they are missing a huge ingredient in the dictionary, namely they have no Riemann-Roch theorem over F_1. Still, it is a very suggestive similarity, and something to keep an eye on in the future... \"</p>\n" } ]
Question: What is precisely still missing in Connes&#39; approach to RH? Answer 1: <p>First, recall the step's of Weil's proof, other than defining the surface:</p> <ol> <li>Develop an intersection theory of curves on surfaces.</li> <li>Show that the intersection of two specially chosen curves is equal to a coefficient of the zeta function.</li> <li>Prove the Hodge index theorem using the Riemann-Roch theorem for surfaces (or is there another proof?).</li> <li>By playing around with the two curves from step 2 and a few other simple curves, deduce an inequality for each coefficient of the logarithmic derivative of the zeta function.</li> <li>Conclude the Riemann hypothesis for the roots of the zeta function.</li> </ol> <p>Any approach which closely follows Weil's strategy would have to find analogues of these 5 steps.</p> <p>I think most people would say the crux of the issue is step 3 - it would be very interesting if some abstract/categorical/algebraic method produced any inequality involving the Riemann zeta function or related functions that was not previously known, even if it was short of the full Riemann hypothesis. Equally, many people fear that if we generalize too far from the world of curves and surfaces that we know and love, we might lose the ability to prove concrete inequalities like this. nfdc23 discussed this issue in the comments.</p> <hr> <p>In the survey article you linke, Connes says:</p> <blockquote> <p>At this point, what is missing is an intersection theory and a Riemann-Roch theorem on the square of the arithmetic site.</p> </blockquote> <p>I think this is still the case - the Riemann-Roch theorem in the paper you link seems to be for certain curves in the scaling site. (If it were for the surface, I wouldn't expect to see it on the arXiv until the Riemann hypothesis were deduced from it, or a very convincing argument was found that it was not sufficient to prove the Riemann hypothesis!)</p> <hr> <p>I'm not sure how helpful knowledge of toposes (alone) will be for understanding this issue. The reason is that the key parts of any definition of intersection theory of curves on surfaces do not involve very much the topos structure of the surface, but rather other structures. </p> <p>I know of a few different perspectives on intersection theory:</p> <ol> <li><p>Add ample classes, apply a moving lemma, and then concretely count the intersections of curves. Here the main issues are global geometric considerations - ample divisors and their relationships to projective embeddings and hyperplanes and so on. I don't think toposes are a very useful tool for understanding these things.</p></li> <li><p>Apply local formulas for the intersection number, and then sum over the point of intersection. This is more promising as e.g. Serre's formula is cohomological in nature, but it's the wrong kind of cohomology - for modules, not for toposes. I think it might be hard to define the right cohomology groups over semirings.</p></li> <li><p>Take the cup product of their classes in a suitable cohomology theory. This at first looks promising, because one of our possible choices for a cohomology theory, etale cohomology, is defined using a topos. Unfortunately, because etale cohomology is naturally l-adic, it is difficult to establish positivity from it. Of course one could look to generalizing Deligne's proof of the Riemann hypothesis, but that is a different matter and much more complicated. (Peter Sarnak has suggested that Deligne's proof provides a guide to how one should try to prove the Riemann hypothesis for number fields, but it is not the etale topos but rather the use of fmailies in the proof that he wants to mimic.)</p></li> <li><p>View one curve as an honest curve and the other as the divisor of a line bundle and pull the line bundle back to the curve, then compute its degree. This is, I think, the easiest way to define the intersection number, but one needs some of the other characterizations to prove its properties (even symmetry). This looks to me like it is at least potentially meaningful in some topos and I don't know exactly to what extent it can be applied to the arithmetic site.</p></li> </ol> <p>Remember that even among locally ringed spaces, the ones where you have a good intersection theory are very specific (If you want to intersect arbitrary closed subsets I think you need smooth proper varieties over a field). For locally semiringed topoi, the situation is presumably much worse.</p> <p>No one is saying it is impossible to develop an applicable intersection theory, but I don't think anyone really knows how to begin -- if they did, I'm sure they would be working furiously on it.</p> Answer 2: <p>There is a very interesting book which appeared recently precisely on this topic: <a href="https://books.google.com/books?id=RbhEAgAAQBAJ&amp;printsec=frontcover&amp;source=gbs_ge_summary_r&amp;cad=0#v=onepage&amp;q&amp;f=false" rel="noreferrer">"The Riemann Hypothesis for Function Fields"</a> by Machiel van Frankenhuijsen. His attempt at answering the OP's question about what is missing is in the last two chapters.</p> <p>My impression is that Connes' work on RH, like that of great painters such as Picasso, can be divided into different periods. I would say the first period is about exploring connections to statistical mechanics and the notion of symmetry breaking. The main result is his article with Bost on the now called Bost-Connes system. However, this did not yet contain an approach towards RH. The second period is that of his Selecta paper with the trace formula. Finally the third period involves having "fun with $\mathbb{F}_1$", hyperreals, toposes, etc. The third period is about constructing suitable objects which would allow one to mimic Weil's proof. I am not familiar with this last development, but I think Will did a good job at explaining what is missing regarding the approach from this third period. The book I mentioned and the little I know concern the second period.</p> <p>Let me first give a general somewhat vague opinion: if one has a reasonably clear description of what $A$ and $B$ are, then it is easier to prove a conjecture which says $A=B$ than one which says $A\neq B$ (or $A&gt;B$, $A\ge B$, etc.). For example, I think the <a href="https://arxiv.org/abs/1508.07205" rel="noreferrer">recent work</a> by Kronheimer and Mrowka is significant progress on the four-color theorem. The latter says $A&gt;B$ where $B=0$ and $A$ is the number of Tait colorings. They produced some invariant $C$ which they showed satisfies $C&gt;B$ and reduced the problem to the conjecture $A=C$. Connes' Selecta article reduced Weil positivity (an $A\ge B$ type problem) and thus RH to an $A=B$ type problem, i.e., proving his global trace formula which is formally similar to a theorem of Guillemin and Sternberg for flows on Riemannian manifolds. He also proved the semi-local analogue of his formula. Take a finite group $G$ with a unitary representation $U$ and consider the matrix ${\rm tr}(U(gh^{-1}))$ with rows and columns indexed by $g,h\in G$. This matrix is positive semidefinite. Weil positivity follows from Connes' global trace formula for essentially the same reason. </p> <p>Proving the trace formula, prima facie, looks like a problem of harmonic analysis or even constructive quantum field theory since one has a formally precise but still physicsy looking infinite-dimensional trace to make sense of by introducing UV and IR cutoffs and then analyzing the limit of removing them. There has been follow-ups by Haran (<a href="https://books.google.com/books?hl=en&amp;lr=&amp;id=eGIbCAAAQBAJ&amp;oi=fnd&amp;pg=PA93&amp;ots=YkfLo5jT04&amp;sig=lqQNwqZUfh_obuH8YVvY4IeYvrg&amp;redir_esc=y#v=onepage&amp;q&amp;f=false" rel="noreferrer">here</a> and <a href="https://books.google.com/books?id=bsUkkEraBvAC&amp;printsec=frontcover#v=onepage&amp;q&amp;f=false" rel="noreferrer">here</a>) and Burnol (<a href="https://arxiv.org/abs/math/0101068" rel="noreferrer">here</a> and <a href="https://arxiv.org/abs/math/0112254" rel="noreferrer">here</a>) but, as far as I know, the trace formula has not been proved in this approach, even in the function field case where at least we know it should be true.</p> Answer 3: <p>Terence Tao answered this question as follows:</p> <p>" He (Connes) also discussed his recent paper with Consani in which they managed to view the (completed) Riemann zeta function as the Hasse-Weil zeta function of a curve over F_1, where F_1 is interpreted as the tropical ring {0,1}. So the Riemann hypothesis for the integers is now expressed in a very similar fashion to the Riemann hypothesis for curves, which has a number of proofs. Unfortunately, they are missing a huge ingredient in the dictionary, namely they have no Riemann-Roch theorem over F_1. Still, it is a very suggestive similarity, and something to keep an eye on in the future... "</p> Answer 4: <p>First, recall the step's of Weil's proof, other than defining the surface:</p> <ol> <li>Develop an intersection theory of curves on surfaces.</li> <li>Show that the intersection of two specially chosen curves is equal to a coefficient of the zeta function.</li> <li>Prove the Hodge index theorem using the Riemann-Roch theorem for surfaces (or is there another proof?).</li> <li>By playing around with the two curves from step 2 and a few other simple curves, deduce an inequality for each coefficient of the logarithmic derivative of the zeta function.</li> <li>Conclude the Riemann hypothesis for the roots of the zeta function.</li> </ol> <p>Any approach which closely follows Weil's strategy would have to find analogues of these 5 steps.</p> <p>I think most people would say the crux of the issue is step 3 - it would be very interesting if some abstract/categorical/algebraic method produced any inequality involving the Riemann zeta function or related functions that was not previously known, even if it was short of the full Riemann hypothesis. Equally, many people fear that if we generalize too far from the world of curves and surfaces that we know and love, we might lose the ability to prove concrete inequalities like this. nfdc23 discussed this issue in the comments.</p> <hr> <p>In the survey article you linke, Connes says:</p> <blockquote> <p>At this point, what is missing is an intersection theory and a Riemann-Roch theorem on the square of the arithmetic site.</p> </blockquote> <p>I think this is still the case - the Riemann-Roch theorem in the paper you link seems to be for certain curves in the scaling site. (If it were for the surface, I wouldn't expect to see it on the arXiv until the Riemann hypothesis were deduced from it, or a very convincing argument was found that it was not sufficient to prove the Riemann hypothesis!)</p> <hr> <p>I'm not sure how helpful knowledge of toposes (alone) will be for understanding this issue. The reason is that the key parts of any definition of intersection theory of curves on surfaces do not involve very much the topos structure of the surface, but rather other structures. </p> <p>I know of a few different perspectives on intersection theory:</p> <ol> <li><p>Add ample classes, apply a moving lemma, and then concretely count the intersections of curves. Here the main issues are global geometric considerations - ample divisors and their relationships to projective embeddings and hyperplanes and so on. I don't think toposes are a very useful tool for understanding these things.</p></li> <li><p>Apply local formulas for the intersection number, and then sum over the point of intersection. This is more promising as e.g. Serre's formula is cohomological in nature, but it's the wrong kind of cohomology - for modules, not for toposes. I think it might be hard to define the right cohomology groups over semirings.</p></li> <li><p>Take the cup product of their classes in a suitable cohomology theory. This at first looks promising, because one of our possible choices for a cohomology theory, etale cohomology, is defined using a topos. Unfortunately, because etale cohomology is naturally l-adic, it is difficult to establish positivity from it. Of course one could look to generalizing Deligne's proof of the Riemann hypothesis, but that is a different matter and much more complicated. (Peter Sarnak has suggested that Deligne's proof provides a guide to how one should try to prove the Riemann hypothesis for number fields, but it is not the etale topos but rather the use of fmailies in the proof that he wants to mimic.)</p></li> <li><p>View one curve as an honest curve and the other as the divisor of a line bundle and pull the line bundle back to the curve, then compute its degree. This is, I think, the easiest way to define the intersection number, but one needs some of the other characterizations to prove its properties (even symmetry). This looks to me like it is at least potentially meaningful in some topos and I don't know exactly to what extent it can be applied to the arithmetic site.</p></li> </ol> <p>Remember that even among locally ringed spaces, the ones where you have a good intersection theory are very specific (If you want to intersect arbitrary closed subsets I think you need smooth proper varieties over a field). For locally semiringed topoi, the situation is presumably much worse.</p> <p>No one is saying it is impossible to develop an applicable intersection theory, but I don't think anyone really knows how to begin -- if they did, I'm sure they would be working furiously on it.</p> Answer 5: <p>There is a very interesting book which appeared recently precisely on this topic: <a href="https://books.google.com/books?id=RbhEAgAAQBAJ&amp;printsec=frontcover&amp;source=gbs_ge_summary_r&amp;cad=0#v=onepage&amp;q&amp;f=false" rel="noreferrer">"The Riemann Hypothesis for Function Fields"</a> by Machiel van Frankenhuijsen. His attempt at answering the OP's question about what is missing is in the last two chapters.</p> <p>My impression is that Connes' work on RH, like that of great painters such as Picasso, can be divided into different periods. I would say the first period is about exploring connections to statistical mechanics and the notion of symmetry breaking. The main result is his article with Bost on the now called Bost-Connes system. However, this did not yet contain an approach towards RH. The second period is that of his Selecta paper with the trace formula. Finally the third period involves having "fun with $\mathbb{F}_1$", hyperreals, toposes, etc. The third period is about constructing suitable objects which would allow one to mimic Weil's proof. I am not familiar with this last development, but I think Will did a good job at explaining what is missing regarding the approach from this third period. The book I mentioned and the little I know concern the second period.</p> <p>Let me first give a general somewhat vague opinion: if one has a reasonably clear description of what $A$ and $B$ are, then it is easier to prove a conjecture which says $A=B$ than one which says $A\neq B$ (or $A&gt;B$, $A\ge B$, etc.). For example, I think the <a href="https://arxiv.org/abs/1508.07205" rel="noreferrer">recent work</a> by Kronheimer and Mrowka is significant progress on the four-color theorem. The latter says $A&gt;B$ where $B=0$ and $A$ is the number of Tait colorings. They produced some invariant $C$ which they showed satisfies $C&gt;B$ and reduced the problem to the conjecture $A=C$. Connes' Selecta article reduced Weil positivity (an $A\ge B$ type problem) and thus RH to an $A=B$ type problem, i.e., proving his global trace formula which is formally similar to a theorem of Guillemin and Sternberg for flows on Riemannian manifolds. He also proved the semi-local analogue of his formula. Take a finite group $G$ with a unitary representation $U$ and consider the matrix ${\rm tr}(U(gh^{-1}))$ with rows and columns indexed by $g,h\in G$. This matrix is positive semidefinite. Weil positivity follows from Connes' global trace formula for essentially the same reason. </p> <p>Proving the trace formula, prima facie, looks like a problem of harmonic analysis or even constructive quantum field theory since one has a formally precise but still physicsy looking infinite-dimensional trace to make sense of by introducing UV and IR cutoffs and then analyzing the limit of removing them. There has been follow-ups by Haran (<a href="https://books.google.com/books?hl=en&amp;lr=&amp;id=eGIbCAAAQBAJ&amp;oi=fnd&amp;pg=PA93&amp;ots=YkfLo5jT04&amp;sig=lqQNwqZUfh_obuH8YVvY4IeYvrg&amp;redir_esc=y#v=onepage&amp;q&amp;f=false" rel="noreferrer">here</a> and <a href="https://books.google.com/books?id=bsUkkEraBvAC&amp;printsec=frontcover#v=onepage&amp;q&amp;f=false" rel="noreferrer">here</a>) and Burnol (<a href="https://arxiv.org/abs/math/0101068" rel="noreferrer">here</a> and <a href="https://arxiv.org/abs/math/0112254" rel="noreferrer">here</a>) but, as far as I know, the trace formula has not been proved in this approach, even in the function field case where at least we know it should be true.</p> Answer 6: <p>Terence Tao answered this question as follows:</p> <p>" He (Connes) also discussed his recent paper with Consani in which they managed to view the (completed) Riemann zeta function as the Hasse-Weil zeta function of a curve over F_1, where F_1 is interpreted as the tropical ring {0,1}. So the Riemann hypothesis for the integers is now expressed in a very similar fashion to the Riemann hypothesis for curves, which has a number of proofs. Unfortunately, they are missing a huge ingredient in the dictionary, namely they have no Riemann-Roch theorem over F_1. Still, it is a very suggestive similarity, and something to keep an eye on in the future... "</p> Answer 7: <p>First, recall the step's of Weil's proof, other than defining the surface:</p> <ol> <li>Develop an intersection theory of curves on surfaces.</li> <li>Show that the intersection of two specially chosen curves is equal to a coefficient of the zeta function.</li> <li>Prove the Hodge index theorem using the Riemann-Roch theorem for surfaces (or is there another proof?).</li> <li>By playing around with the two curves from step 2 and a few other simple curves, deduce an inequality for each coefficient of the logarithmic derivative of the zeta function.</li> <li>Conclude the Riemann hypothesis for the roots of the zeta function.</li> </ol> <p>Any approach which closely follows Weil's strategy would have to find analogues of these 5 steps.</p> <p>I think most people would say the crux of the issue is step 3 - it would be very interesting if some abstract/categorical/algebraic method produced any inequality involving the Riemann zeta function or related functions that was not previously known, even if it was short of the full Riemann hypothesis. Equally, many people fear that if we generalize too far from the world of curves and surfaces that we know and love, we might lose the ability to prove concrete inequalities like this. nfdc23 discussed this issue in the comments.</p> <hr> <p>In the survey article you linke, Connes says:</p> <blockquote> <p>At this point, what is missing is an intersection theory and a Riemann-Roch theorem on the square of the arithmetic site.</p> </blockquote> <p>I think this is still the case - the Riemann-Roch theorem in the paper you link seems to be for certain curves in the scaling site. (If it were for the surface, I wouldn't expect to see it on the arXiv until the Riemann hypothesis were deduced from it, or a very convincing argument was found that it was not sufficient to prove the Riemann hypothesis!)</p> <hr> <p>I'm not sure how helpful knowledge of toposes (alone) will be for understanding this issue. The reason is that the key parts of any definition of intersection theory of curves on surfaces do not involve very much the topos structure of the surface, but rather other structures. </p> <p>I know of a few different perspectives on intersection theory:</p> <ol> <li><p>Add ample classes, apply a moving lemma, and then concretely count the intersections of curves. Here the main issues are global geometric considerations - ample divisors and their relationships to projective embeddings and hyperplanes and so on. I don't think toposes are a very useful tool for understanding these things.</p></li> <li><p>Apply local formulas for the intersection number, and then sum over the point of intersection. This is more promising as e.g. Serre's formula is cohomological in nature, but it's the wrong kind of cohomology - for modules, not for toposes. I think it might be hard to define the right cohomology groups over semirings.</p></li> <li><p>Take the cup product of their classes in a suitable cohomology theory. This at first looks promising, because one of our possible choices for a cohomology theory, etale cohomology, is defined using a topos. Unfortunately, because etale cohomology is naturally l-adic, it is difficult to establish positivity from it. Of course one could look to generalizing Deligne's proof of the Riemann hypothesis, but that is a different matter and much more complicated. (Peter Sarnak has suggested that Deligne's proof provides a guide to how one should try to prove the Riemann hypothesis for number fields, but it is not the etale topos but rather the use of fmailies in the proof that he wants to mimic.)</p></li> <li><p>View one curve as an honest curve and the other as the divisor of a line bundle and pull the line bundle back to the curve, then compute its degree. This is, I think, the easiest way to define the intersection number, but one needs some of the other characterizations to prove its properties (even symmetry). This looks to me like it is at least potentially meaningful in some topos and I don't know exactly to what extent it can be applied to the arithmetic site.</p></li> </ol> <p>Remember that even among locally ringed spaces, the ones where you have a good intersection theory are very specific (If you want to intersect arbitrary closed subsets I think you need smooth proper varieties over a field). For locally semiringed topoi, the situation is presumably much worse.</p> <p>No one is saying it is impossible to develop an applicable intersection theory, but I don't think anyone really knows how to begin -- if they did, I'm sure they would be working furiously on it.</p> Answer 8: <p>There is a very interesting book which appeared recently precisely on this topic: <a href="https://books.google.com/books?id=RbhEAgAAQBAJ&amp;printsec=frontcover&amp;source=gbs_ge_summary_r&amp;cad=0#v=onepage&amp;q&amp;f=false" rel="noreferrer">"The Riemann Hypothesis for Function Fields"</a> by Machiel van Frankenhuijsen. His attempt at answering the OP's question about what is missing is in the last two chapters.</p> <p>My impression is that Connes' work on RH, like that of great painters such as Picasso, can be divided into different periods. I would say the first period is about exploring connections to statistical mechanics and the notion of symmetry breaking. The main result is his article with Bost on the now called Bost-Connes system. However, this did not yet contain an approach towards RH. The second period is that of his Selecta paper with the trace formula. Finally the third period involves having "fun with $\mathbb{F}_1$", hyperreals, toposes, etc. The third period is about constructing suitable objects which would allow one to mimic Weil's proof. I am not familiar with this last development, but I think Will did a good job at explaining what is missing regarding the approach from this third period. The book I mentioned and the little I know concern the second period.</p> <p>Let me first give a general somewhat vague opinion: if one has a reasonably clear description of what $A$ and $B$ are, then it is easier to prove a conjecture which says $A=B$ than one which says $A\neq B$ (or $A&gt;B$, $A\ge B$, etc.). For example, I think the <a href="https://arxiv.org/abs/1508.07205" rel="noreferrer">recent work</a> by Kronheimer and Mrowka is significant progress on the four-color theorem. The latter says $A&gt;B$ where $B=0$ and $A$ is the number of Tait colorings. They produced some invariant $C$ which they showed satisfies $C&gt;B$ and reduced the problem to the conjecture $A=C$. Connes' Selecta article reduced Weil positivity (an $A\ge B$ type problem) and thus RH to an $A=B$ type problem, i.e., proving his global trace formula which is formally similar to a theorem of Guillemin and Sternberg for flows on Riemannian manifolds. He also proved the semi-local analogue of his formula. Take a finite group $G$ with a unitary representation $U$ and consider the matrix ${\rm tr}(U(gh^{-1}))$ with rows and columns indexed by $g,h\in G$. This matrix is positive semidefinite. Weil positivity follows from Connes' global trace formula for essentially the same reason. </p> <p>Proving the trace formula, prima facie, looks like a problem of harmonic analysis or even constructive quantum field theory since one has a formally precise but still physicsy looking infinite-dimensional trace to make sense of by introducing UV and IR cutoffs and then analyzing the limit of removing them. There has been follow-ups by Haran (<a href="https://books.google.com/books?hl=en&amp;lr=&amp;id=eGIbCAAAQBAJ&amp;oi=fnd&amp;pg=PA93&amp;ots=YkfLo5jT04&amp;sig=lqQNwqZUfh_obuH8YVvY4IeYvrg&amp;redir_esc=y#v=onepage&amp;q&amp;f=false" rel="noreferrer">here</a> and <a href="https://books.google.com/books?id=bsUkkEraBvAC&amp;printsec=frontcover#v=onepage&amp;q&amp;f=false" rel="noreferrer">here</a>) and Burnol (<a href="https://arxiv.org/abs/math/0101068" rel="noreferrer">here</a> and <a href="https://arxiv.org/abs/math/0112254" rel="noreferrer">here</a>) but, as far as I know, the trace formula has not been proved in this approach, even in the function field case where at least we know it should be true.</p> Answer 9: <p>Terence Tao answered this question as follows:</p> <p>" He (Connes) also discussed his recent paper with Consani in which they managed to view the (completed) Riemann zeta function as the Hasse-Weil zeta function of a curve over F_1, where F_1 is interpreted as the tropical ring {0,1}. So the Riemann hypothesis for the integers is now expressed in a very similar fashion to the Riemann hypothesis for curves, which has a number of proofs. Unfortunately, they are missing a huge ingredient in the dictionary, namely they have no Riemann-Roch theorem over F_1. Still, it is a very suggestive similarity, and something to keep an eye on in the future... "</p> Answer 10: <p>First, recall the step's of Weil's proof, other than defining the surface:</p> <ol> <li>Develop an intersection theory of curves on surfaces.</li> <li>Show that the intersection of two specially chosen curves is equal to a coefficient of the zeta function.</li> <li>Prove the Hodge index theorem using the Riemann-Roch theorem for surfaces (or is there another proof?).</li> <li>By playing around with the two curves from step 2 and a few other simple curves, deduce an inequality for each coefficient of the logarithmic derivative of the zeta function.</li> <li>Conclude the Riemann hypothesis for the roots of the zeta function.</li> </ol> <p>Any approach which closely follows Weil's strategy would have to find analogues of these 5 steps.</p> <p>I think most people would say the crux of the issue is step 3 - it would be very interesting if some abstract/categorical/algebraic method produced any inequality involving the Riemann zeta function or related functions that was not previously known, even if it was short of the full Riemann hypothesis. Equally, many people fear that if we generalize too far from the world of curves and surfaces that we know and love, we might lose the ability to prove concrete inequalities like this. nfdc23 discussed this issue in the comments.</p> <hr> <p>In the survey article you linke, Connes says:</p> <blockquote> <p>At this point, what is missing is an intersection theory and a Riemann-Roch theorem on the square of the arithmetic site.</p> </blockquote> <p>I think this is still the case - the Riemann-Roch theorem in the paper you link seems to be for certain curves in the scaling site. (If it were for the surface, I wouldn't expect to see it on the arXiv until the Riemann hypothesis were deduced from it, or a very convincing argument was found that it was not sufficient to prove the Riemann hypothesis!)</p> <hr> <p>I'm not sure how helpful knowledge of toposes (alone) will be for understanding this issue. The reason is that the key parts of any definition of intersection theory of curves on surfaces do not involve very much the topos structure of the surface, but rather other structures. </p> <p>I know of a few different perspectives on intersection theory:</p> <ol> <li><p>Add ample classes, apply a moving lemma, and then concretely count the intersections of curves. Here the main issues are global geometric considerations - ample divisors and their relationships to projective embeddings and hyperplanes and so on. I don't think toposes are a very useful tool for understanding these things.</p></li> <li><p>Apply local formulas for the intersection number, and then sum over the point of intersection. This is more promising as e.g. Serre's formula is cohomological in nature, but it's the wrong kind of cohomology - for modules, not for toposes. I think it might be hard to define the right cohomology groups over semirings.</p></li> <li><p>Take the cup product of their classes in a suitable cohomology theory. This at first looks promising, because one of our possible choices for a cohomology theory, etale cohomology, is defined using a topos. Unfortunately, because etale cohomology is naturally l-adic, it is difficult to establish positivity from it. Of course one could look to generalizing Deligne's proof of the Riemann hypothesis, but that is a different matter and much more complicated. (Peter Sarnak has suggested that Deligne's proof provides a guide to how one should try to prove the Riemann hypothesis for number fields, but it is not the etale topos but rather the use of fmailies in the proof that he wants to mimic.)</p></li> <li><p>View one curve as an honest curve and the other as the divisor of a line bundle and pull the line bundle back to the curve, then compute its degree. This is, I think, the easiest way to define the intersection number, but one needs some of the other characterizations to prove its properties (even symmetry). This looks to me like it is at least potentially meaningful in some topos and I don't know exactly to what extent it can be applied to the arithmetic site.</p></li> </ol> <p>Remember that even among locally ringed spaces, the ones where you have a good intersection theory are very specific (If you want to intersect arbitrary closed subsets I think you need smooth proper varieties over a field). For locally semiringed topoi, the situation is presumably much worse.</p> <p>No one is saying it is impossible to develop an applicable intersection theory, but I don't think anyone really knows how to begin -- if they did, I'm sure they would be working furiously on it.</p> Answer 11: <p>There is a very interesting book which appeared recently precisely on this topic: <a href="https://books.google.com/books?id=RbhEAgAAQBAJ&amp;printsec=frontcover&amp;source=gbs_ge_summary_r&amp;cad=0#v=onepage&amp;q&amp;f=false" rel="noreferrer">"The Riemann Hypothesis for Function Fields"</a> by Machiel van Frankenhuijsen. His attempt at answering the OP's question about what is missing is in the last two chapters.</p> <p>My impression is that Connes' work on RH, like that of great painters such as Picasso, can be divided into different periods. I would say the first period is about exploring connections to statistical mechanics and the notion of symmetry breaking. The main result is his article with Bost on the now called Bost-Connes system. However, this did not yet contain an approach towards RH. The second period is that of his Selecta paper with the trace formula. Finally the third period involves having "fun with $\mathbb{F}_1$", hyperreals, toposes, etc. The third period is about constructing suitable objects which would allow one to mimic Weil's proof. I am not familiar with this last development, but I think Will did a good job at explaining what is missing regarding the approach from this third period. The book I mentioned and the little I know concern the second period.</p> <p>Let me first give a general somewhat vague opinion: if one has a reasonably clear description of what $A$ and $B$ are, then it is easier to prove a conjecture which says $A=B$ than one which says $A\neq B$ (or $A&gt;B$, $A\ge B$, etc.). For example, I think the <a href="https://arxiv.org/abs/1508.07205" rel="noreferrer">recent work</a> by Kronheimer and Mrowka is significant progress on the four-color theorem. The latter says $A&gt;B$ where $B=0$ and $A$ is the number of Tait colorings. They produced some invariant $C$ which they showed satisfies $C&gt;B$ and reduced the problem to the conjecture $A=C$. Connes' Selecta article reduced Weil positivity (an $A\ge B$ type problem) and thus RH to an $A=B$ type problem, i.e., proving his global trace formula which is formally similar to a theorem of Guillemin and Sternberg for flows on Riemannian manifolds. He also proved the semi-local analogue of his formula. Take a finite group $G$ with a unitary representation $U$ and consider the matrix ${\rm tr}(U(gh^{-1}))$ with rows and columns indexed by $g,h\in G$. This matrix is positive semidefinite. Weil positivity follows from Connes' global trace formula for essentially the same reason. </p> <p>Proving the trace formula, prima facie, looks like a problem of harmonic analysis or even constructive quantum field theory since one has a formally precise but still physicsy looking infinite-dimensional trace to make sense of by introducing UV and IR cutoffs and then analyzing the limit of removing them. There has been follow-ups by Haran (<a href="https://books.google.com/books?hl=en&amp;lr=&amp;id=eGIbCAAAQBAJ&amp;oi=fnd&amp;pg=PA93&amp;ots=YkfLo5jT04&amp;sig=lqQNwqZUfh_obuH8YVvY4IeYvrg&amp;redir_esc=y#v=onepage&amp;q&amp;f=false" rel="noreferrer">here</a> and <a href="https://books.google.com/books?id=bsUkkEraBvAC&amp;printsec=frontcover#v=onepage&amp;q&amp;f=false" rel="noreferrer">here</a>) and Burnol (<a href="https://arxiv.org/abs/math/0101068" rel="noreferrer">here</a> and <a href="https://arxiv.org/abs/math/0112254" rel="noreferrer">here</a>) but, as far as I know, the trace formula has not been proved in this approach, even in the function field case where at least we know it should be true.</p> Answer 12: <p>Terence Tao answered this question as follows:</p> <p>" He (Connes) also discussed his recent paper with Consani in which they managed to view the (completed) Riemann zeta function as the Hasse-Weil zeta function of a curve over F_1, where F_1 is interpreted as the tropical ring {0,1}. So the Riemann hypothesis for the integers is now expressed in a very similar fashion to the Riemann hypothesis for curves, which has a number of proofs. Unfortunately, they are missing a huge ingredient in the dictionary, namely they have no Riemann-Roch theorem over F_1. Still, it is a very suggestive similarity, and something to keep an eye on in the future... "</p>
Unusual symmetries of the Cayley-Menger determinant for the volume of tetrahedra
259,664
54
https://mathoverflow.net/questions/259664/unusual-symmetries-of-the-cayley-menger-determinant-for-the-volume-of-tetrahedra
<p>Suppose you have a tetrahedron <span class="math-container">$T$</span> in Euclidean space with edge lengths <span class="math-container">$\ell_{01}$</span>, <span class="math-container">$\ell_{02}$</span>, <span class="math-container">$\ell_{03}$</span>, <span class="math-container">$\ell_{12}$</span>, <span class="math-container">$\ell_{13}$</span>, and <span class="math-container">$\ell_{23}$</span>. Now consider the tetrahedron <span class="math-container">$T'$</span> with edge lengths <span class="math-container">$$\begin{aligned} \ell'_{02} &amp;= \ell_{02} &amp; \ell'_{13} &amp;= \ell_{13}\\ \ell'_{01} &amp;= s-\ell_{01} &amp; \ell'_{12} &amp;= s-\ell_{12}\\ \ell'_{23} &amp;= s-\ell_{23}&amp; \ell'_{03} &amp;= s-\ell_{03} \end{aligned} $$</span> where <span class="math-container">$s = (\ell_{01} + \ell_{12} + \ell_{23} + \ell_{03})/2$</span>. If the edge lengths of <span class="math-container">$T'$</span> are positive and satisfy the triangle inequality, then the volume of <span class="math-container">$T'$</span> equals the volume of <span class="math-container">$T$</span>. In particular, if <span class="math-container">$T$</span> is a flat tetrahedron in <span class="math-container">$\mathbb{R}^2$</span>, then <span class="math-container">$T'$</span> is as well. This is easily verified by plugging the values <span class="math-container">$\ell'_{ij}$</span> above into the Cayley-Menger determinant.</p> <p>In fact, it's possible to show that the linear symmetries of <span class="math-container">$\mathbb{R}^6$</span> that preserve the Cayley-Menger determinant form the Weyl group <span class="math-container">$D_6$</span>, of order <span class="math-container">$2^5 * 6! = 23040$</span>. This is a factor of <span class="math-container">$15$</span> times larger than the natural geometric symmetries obtained by permuting the vertices of the tetrahedron and negating the coordinates.</p> <p>The transformations don't always take Euclidean tetrahedra to Euclidean tetrahedra, but they do sometimes. For instance, if you start with an equilateral tetrahedron <span class="math-container">$T$</span> with all side lengths equal to <span class="math-container">$1$</span>, then <span class="math-container">$T'$</span> is also an equilateral tetrahedron. Thus if <span class="math-container">$T$</span> is a generic Euclidean tetrahedron close to equilateral, <span class="math-container">$T'$</span> will also be one, and <span class="math-container">$T$</span> and <span class="math-container">$T'$</span> will not be related by a Euclidean symmetry.</p> <p>I can't be the first person to observe this. (In fact, I vaguely recall hearing about this in the context of quantum groups and the Jones polynomial.) What's the history? How to best understand these transformations (without expanding out the determinant)? Are <span class="math-container">$T$</span> and <span class="math-container">$T'$</span> scissors congruent? Etc.</p>
[ { "answer_id": 259680, "score": 39, "text": "<p>These are the so-called Regge symmetries, described by T. Regge in a 1970-ish paper. For a bit on it, with references, see the paper</p>\n\n<p><em>Philip P. Boalch</em>, MR 2342290 <a href=\"http://dx.doi.org/10.1007/s00220-007-0328-x\"><strong>Regge and Okamoto symmetries</strong></a>, <em>Comm. Math. Phys.</em> <strong>276</strong> (2007), no. 1, 117--130.\\</p>\n" }, { "answer_id": 259767, "score": 21, "text": "<p>To address the scissors congruence question at the end of the post: the Regge symmetries produce tetrahedra which are scissors congruent. This is proved in Section 6 (Theorem 9 and Corollary 10) of </p>\n\n<ul>\n<li>J. Roberts. Classical <span class=\"math-container\">$6j$</span>-symbols and the tetrahedron. Geom. &amp; Top. 3 (1999), pp. 21-66. <a href=\"http://arxiv.org/abs/math-ph/9812013\" rel=\"nofollow noreferrer\">(link to paper on arXiv)</a></li>\n</ul>\n\n<p>The argument is indirect, proving that the Regge symmetries preserve volume and Dehn invariant, then deducing the statement on scissors congruence from Sydler's theorem.</p>\n\n<p>The corresponding statement for hyperbolic tetrahedra was proved (by construction) in </p>\n\n<ul>\n<li>Y. Mohanty. The Regge symmetry is a scissors congruence in hyperbolic space. Alg. Geom. Top. 3 (2003), 1-31. <a href=\"https://arxiv.org/abs/math/0301318\" rel=\"nofollow noreferrer\">(link to paper on arXiv)</a></li>\n</ul>\n" }, { "answer_id": 259989, "score": 19, "text": "<p>Let me give a geometric interpretation for the case of tetrahedra of volume zero. The statement becomes as follows:</p>\n\n<p>Given four positive numbers $a,b,c,d$ that satisfy the quadrangle inequalities, denote\n$$\ns=\\frac{a+b+c+d}2, \\quad a'=s-a \\text{ etc.}\n$$\nTake a quadrilateral with side lengths $a,b,c,d$ (in this cyclic order). If its diagonals have lengths $x,y$, then there exists a quadrilateral with side lengths $a',b',c',d'$ and the same diagonal lengths.</p>\n\n<p>Again, this can be proved by a direct computation, but Arseniy Akopyan noticed that this is equivalent to the Ivory lemma: diagonals in a curvilinear quadrilateral formed by four confocal conics are equal, see the left half of the picture.</p>\n\n<p><a href=\"https://i.stack.imgur.com/7KCQP.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/7KCQP.png\" alt=\"enter image description here\"></a></p>\n\n<p>Indeed, with two opposite vertices of an $(a,b,c,d)$-quadrilateral as the foci, draw two ellipses and two hyperbolas through the other two vertices. This gives a curvilinear quadrilateral whose diagonal is also a diagonal of our $(a,b,c,d)$-quadrilateral, see the right half of the picture. The relations between $a,b,c,d$ and $a',b',c',d'$:\n$$\na+b=c'+d', c-d=d'-c' \\text{ etc.}\n$$\nimply that the other two vertices of this curvilinear quadrilateral form together with the foci an $(a',b',c',d')$-quadrilateral.</p>\n\n<p>As a side remark: I studied configuration spaces of planar quadrilaterals in <a href=\"https://arxiv.org/abs/1501.07157\" rel=\"noreferrer\" title=\"this preprint\">this preprint</a>, but didn't know about Regge symmetries. Thanks to Dylan and Igor for bringing this up!</p>\n\n<p>There is an analog of the Ivory lemma in dimension three, but there is no word about volumes in it. Still, it would be interesting to look at the corresponding configuration of confocal surfaces (one of the families degenerates).</p>\n\n<p>Also note that the Ivory lemma holds in the hyperbolic and spherical geometry as well.</p>\n" }, { "answer_id": 259847, "score": 7, "text": "<p>It may be worth noting that for some choices of six lengths that there exist as many as 30 inequivalent tetrahedra with these six lengths. </p>\n" }, { "answer_id": 376730, "score": 4, "text": "<p>The following comment in the question intrigued me:</p>\n<blockquote>\n<p>In fact, it's possible to show that the linear symmetries of\n<span class=\"math-container\">$\\mathbb{R}^6$</span> that preserve the Cayley-Menger determinant form the\nWeyl group <span class=\"math-container\">$D_6$</span>, of order <span class=\"math-container\">$2^5 * 6! = 23040$</span>.</p>\n</blockquote>\n<p>In particular, this suggests that it should be possible to perform a change of basis such that the Cayley–Menger determinant is a symmetric polynomial in the six variables, which is also invariant under even numbers of sign changes.</p>\n<p>Some experimentation confirmed that the following variables work in the sense that the geometric and Regge symmetries all preserve the 12-element set <span class=\"math-container\">$\\{ \\pm u, \\pm v, \\pm w, \\pm x, \\pm y, \\pm z \\}$</span>:</p>\n<ul>\n<li><span class=\"math-container\">$u = \\ell_{01} + \\ell_{23}$</span>;</li>\n<li><span class=\"math-container\">$v = \\ell_{02} + \\ell_{13}$</span>;</li>\n<li><span class=\"math-container\">$w = \\ell_{03} + \\ell_{12}$</span>;</li>\n<li><span class=\"math-container\">$x = \\ell_{01} - \\ell_{23}$</span>;</li>\n<li><span class=\"math-container\">$y = \\ell_{02} - \\ell_{13}$</span>;</li>\n<li><span class=\"math-container\">$z = \\ell_{03} - \\ell_{12}$</span>;</li>\n</ul>\n<p>More succinctly, the variables are the three sums of opposite edges and the three differences of opposite edges. The Cayley–Menger determinant can then be shown to be equal to the following 57-term polynomial:</p>\n<p><span class=\"math-container\">$$ uvwxyz - \\dfrac{1}{16} \\sum_\\text{sym}^{6} u^6 + \\dfrac{1}{16} \\sum_\\text{sym}^{30} u^4 v^2 - \\dfrac{1}{8} \\sum_\\text{sym}^{20} u^2 v^2 w^2 $$</span></p>\n<p>where the ‘symmetric sums’ are over all distinct monomials obtained by permuting the six variables (and the number above the sum is the number of terms). This is visibly invariant under the Weyl group <span class=\"math-container\">$D_6$</span> generated by arbitrary permutations and even numbers of sign changes.</p>\n" }, { "answer_id": 259680, "score": 39, "text": "<p>These are the so-called Regge symmetries, described by T. Regge in a 1970-ish paper. For a bit on it, with references, see the paper</p>\n\n<p><em>Philip P. Boalch</em>, MR 2342290 <a href=\"http://dx.doi.org/10.1007/s00220-007-0328-x\"><strong>Regge and Okamoto symmetries</strong></a>, <em>Comm. Math. Phys.</em> <strong>276</strong> (2007), no. 1, 117--130.\\</p>\n" }, { "answer_id": 259767, "score": 21, "text": "<p>To address the scissors congruence question at the end of the post: the Regge symmetries produce tetrahedra which are scissors congruent. This is proved in Section 6 (Theorem 9 and Corollary 10) of </p>\n\n<ul>\n<li>J. Roberts. Classical <span class=\"math-container\">$6j$</span>-symbols and the tetrahedron. Geom. &amp; Top. 3 (1999), pp. 21-66. <a href=\"http://arxiv.org/abs/math-ph/9812013\" rel=\"nofollow noreferrer\">(link to paper on arXiv)</a></li>\n</ul>\n\n<p>The argument is indirect, proving that the Regge symmetries preserve volume and Dehn invariant, then deducing the statement on scissors congruence from Sydler's theorem.</p>\n\n<p>The corresponding statement for hyperbolic tetrahedra was proved (by construction) in </p>\n\n<ul>\n<li>Y. Mohanty. The Regge symmetry is a scissors congruence in hyperbolic space. Alg. Geom. Top. 3 (2003), 1-31. <a href=\"https://arxiv.org/abs/math/0301318\" rel=\"nofollow noreferrer\">(link to paper on arXiv)</a></li>\n</ul>\n" }, { "answer_id": 259989, "score": 19, "text": "<p>Let me give a geometric interpretation for the case of tetrahedra of volume zero. The statement becomes as follows:</p>\n\n<p>Given four positive numbers $a,b,c,d$ that satisfy the quadrangle inequalities, denote\n$$\ns=\\frac{a+b+c+d}2, \\quad a'=s-a \\text{ etc.}\n$$\nTake a quadrilateral with side lengths $a,b,c,d$ (in this cyclic order). If its diagonals have lengths $x,y$, then there exists a quadrilateral with side lengths $a',b',c',d'$ and the same diagonal lengths.</p>\n\n<p>Again, this can be proved by a direct computation, but Arseniy Akopyan noticed that this is equivalent to the Ivory lemma: diagonals in a curvilinear quadrilateral formed by four confocal conics are equal, see the left half of the picture.</p>\n\n<p><a href=\"https://i.stack.imgur.com/7KCQP.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/7KCQP.png\" alt=\"enter image description here\"></a></p>\n\n<p>Indeed, with two opposite vertices of an $(a,b,c,d)$-quadrilateral as the foci, draw two ellipses and two hyperbolas through the other two vertices. This gives a curvilinear quadrilateral whose diagonal is also a diagonal of our $(a,b,c,d)$-quadrilateral, see the right half of the picture. The relations between $a,b,c,d$ and $a',b',c',d'$:\n$$\na+b=c'+d', c-d=d'-c' \\text{ etc.}\n$$\nimply that the other two vertices of this curvilinear quadrilateral form together with the foci an $(a',b',c',d')$-quadrilateral.</p>\n\n<p>As a side remark: I studied configuration spaces of planar quadrilaterals in <a href=\"https://arxiv.org/abs/1501.07157\" rel=\"noreferrer\" title=\"this preprint\">this preprint</a>, but didn't know about Regge symmetries. Thanks to Dylan and Igor for bringing this up!</p>\n\n<p>There is an analog of the Ivory lemma in dimension three, but there is no word about volumes in it. Still, it would be interesting to look at the corresponding configuration of confocal surfaces (one of the families degenerates).</p>\n\n<p>Also note that the Ivory lemma holds in the hyperbolic and spherical geometry as well.</p>\n" }, { "answer_id": 259847, "score": 7, "text": "<p>It may be worth noting that for some choices of six lengths that there exist as many as 30 inequivalent tetrahedra with these six lengths. </p>\n" }, { "answer_id": 376730, "score": 4, "text": "<p>The following comment in the question intrigued me:</p>\n<blockquote>\n<p>In fact, it's possible to show that the linear symmetries of\n<span class=\"math-container\">$\\mathbb{R}^6$</span> that preserve the Cayley-Menger determinant form the\nWeyl group <span class=\"math-container\">$D_6$</span>, of order <span class=\"math-container\">$2^5 * 6! = 23040$</span>.</p>\n</blockquote>\n<p>In particular, this suggests that it should be possible to perform a change of basis such that the Cayley–Menger determinant is a symmetric polynomial in the six variables, which is also invariant under even numbers of sign changes.</p>\n<p>Some experimentation confirmed that the following variables work in the sense that the geometric and Regge symmetries all preserve the 12-element set <span class=\"math-container\">$\\{ \\pm u, \\pm v, \\pm w, \\pm x, \\pm y, \\pm z \\}$</span>:</p>\n<ul>\n<li><span class=\"math-container\">$u = \\ell_{01} + \\ell_{23}$</span>;</li>\n<li><span class=\"math-container\">$v = \\ell_{02} + \\ell_{13}$</span>;</li>\n<li><span class=\"math-container\">$w = \\ell_{03} + \\ell_{12}$</span>;</li>\n<li><span class=\"math-container\">$x = \\ell_{01} - \\ell_{23}$</span>;</li>\n<li><span class=\"math-container\">$y = \\ell_{02} - \\ell_{13}$</span>;</li>\n<li><span class=\"math-container\">$z = \\ell_{03} - \\ell_{12}$</span>;</li>\n</ul>\n<p>More succinctly, the variables are the three sums of opposite edges and the three differences of opposite edges. The Cayley–Menger determinant can then be shown to be equal to the following 57-term polynomial:</p>\n<p><span class=\"math-container\">$$ uvwxyz - \\dfrac{1}{16} \\sum_\\text{sym}^{6} u^6 + \\dfrac{1}{16} \\sum_\\text{sym}^{30} u^4 v^2 - \\dfrac{1}{8} \\sum_\\text{sym}^{20} u^2 v^2 w^2 $$</span></p>\n<p>where the ‘symmetric sums’ are over all distinct monomials obtained by permuting the six variables (and the number above the sum is the number of terms). This is visibly invariant under the Weyl group <span class=\"math-container\">$D_6$</span> generated by arbitrary permutations and even numbers of sign changes.</p>\n" }, { "answer_id": 259680, "score": 39, "text": "<p>These are the so-called Regge symmetries, described by T. Regge in a 1970-ish paper. For a bit on it, with references, see the paper</p>\n\n<p><em>Philip P. Boalch</em>, MR 2342290 <a href=\"http://dx.doi.org/10.1007/s00220-007-0328-x\"><strong>Regge and Okamoto symmetries</strong></a>, <em>Comm. Math. Phys.</em> <strong>276</strong> (2007), no. 1, 117--130.\\</p>\n" }, { "answer_id": 259767, "score": 21, "text": "<p>To address the scissors congruence question at the end of the post: the Regge symmetries produce tetrahedra which are scissors congruent. This is proved in Section 6 (Theorem 9 and Corollary 10) of </p>\n\n<ul>\n<li>J. Roberts. Classical <span class=\"math-container\">$6j$</span>-symbols and the tetrahedron. Geom. &amp; Top. 3 (1999), pp. 21-66. <a href=\"http://arxiv.org/abs/math-ph/9812013\" rel=\"nofollow noreferrer\">(link to paper on arXiv)</a></li>\n</ul>\n\n<p>The argument is indirect, proving that the Regge symmetries preserve volume and Dehn invariant, then deducing the statement on scissors congruence from Sydler's theorem.</p>\n\n<p>The corresponding statement for hyperbolic tetrahedra was proved (by construction) in </p>\n\n<ul>\n<li>Y. Mohanty. The Regge symmetry is a scissors congruence in hyperbolic space. Alg. Geom. Top. 3 (2003), 1-31. <a href=\"https://arxiv.org/abs/math/0301318\" rel=\"nofollow noreferrer\">(link to paper on arXiv)</a></li>\n</ul>\n" }, { "answer_id": 259989, "score": 19, "text": "<p>Let me give a geometric interpretation for the case of tetrahedra of volume zero. The statement becomes as follows:</p>\n\n<p>Given four positive numbers $a,b,c,d$ that satisfy the quadrangle inequalities, denote\n$$\ns=\\frac{a+b+c+d}2, \\quad a'=s-a \\text{ etc.}\n$$\nTake a quadrilateral with side lengths $a,b,c,d$ (in this cyclic order). If its diagonals have lengths $x,y$, then there exists a quadrilateral with side lengths $a',b',c',d'$ and the same diagonal lengths.</p>\n\n<p>Again, this can be proved by a direct computation, but Arseniy Akopyan noticed that this is equivalent to the Ivory lemma: diagonals in a curvilinear quadrilateral formed by four confocal conics are equal, see the left half of the picture.</p>\n\n<p><a href=\"https://i.stack.imgur.com/7KCQP.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/7KCQP.png\" alt=\"enter image description here\"></a></p>\n\n<p>Indeed, with two opposite vertices of an $(a,b,c,d)$-quadrilateral as the foci, draw two ellipses and two hyperbolas through the other two vertices. This gives a curvilinear quadrilateral whose diagonal is also a diagonal of our $(a,b,c,d)$-quadrilateral, see the right half of the picture. The relations between $a,b,c,d$ and $a',b',c',d'$:\n$$\na+b=c'+d', c-d=d'-c' \\text{ etc.}\n$$\nimply that the other two vertices of this curvilinear quadrilateral form together with the foci an $(a',b',c',d')$-quadrilateral.</p>\n\n<p>As a side remark: I studied configuration spaces of planar quadrilaterals in <a href=\"https://arxiv.org/abs/1501.07157\" rel=\"noreferrer\" title=\"this preprint\">this preprint</a>, but didn't know about Regge symmetries. Thanks to Dylan and Igor for bringing this up!</p>\n\n<p>There is an analog of the Ivory lemma in dimension three, but there is no word about volumes in it. Still, it would be interesting to look at the corresponding configuration of confocal surfaces (one of the families degenerates).</p>\n\n<p>Also note that the Ivory lemma holds in the hyperbolic and spherical geometry as well.</p>\n" }, { "answer_id": 259847, "score": 7, "text": "<p>It may be worth noting that for some choices of six lengths that there exist as many as 30 inequivalent tetrahedra with these six lengths. </p>\n" }, { "answer_id": 376730, "score": 4, "text": "<p>The following comment in the question intrigued me:</p>\n<blockquote>\n<p>In fact, it's possible to show that the linear symmetries of\n<span class=\"math-container\">$\\mathbb{R}^6$</span> that preserve the Cayley-Menger determinant form the\nWeyl group <span class=\"math-container\">$D_6$</span>, of order <span class=\"math-container\">$2^5 * 6! = 23040$</span>.</p>\n</blockquote>\n<p>In particular, this suggests that it should be possible to perform a change of basis such that the Cayley–Menger determinant is a symmetric polynomial in the six variables, which is also invariant under even numbers of sign changes.</p>\n<p>Some experimentation confirmed that the following variables work in the sense that the geometric and Regge symmetries all preserve the 12-element set <span class=\"math-container\">$\\{ \\pm u, \\pm v, \\pm w, \\pm x, \\pm y, \\pm z \\}$</span>:</p>\n<ul>\n<li><span class=\"math-container\">$u = \\ell_{01} + \\ell_{23}$</span>;</li>\n<li><span class=\"math-container\">$v = \\ell_{02} + \\ell_{13}$</span>;</li>\n<li><span class=\"math-container\">$w = \\ell_{03} + \\ell_{12}$</span>;</li>\n<li><span class=\"math-container\">$x = \\ell_{01} - \\ell_{23}$</span>;</li>\n<li><span class=\"math-container\">$y = \\ell_{02} - \\ell_{13}$</span>;</li>\n<li><span class=\"math-container\">$z = \\ell_{03} - \\ell_{12}$</span>;</li>\n</ul>\n<p>More succinctly, the variables are the three sums of opposite edges and the three differences of opposite edges. The Cayley–Menger determinant can then be shown to be equal to the following 57-term polynomial:</p>\n<p><span class=\"math-container\">$$ uvwxyz - \\dfrac{1}{16} \\sum_\\text{sym}^{6} u^6 + \\dfrac{1}{16} \\sum_\\text{sym}^{30} u^4 v^2 - \\dfrac{1}{8} \\sum_\\text{sym}^{20} u^2 v^2 w^2 $$</span></p>\n<p>where the ‘symmetric sums’ are over all distinct monomials obtained by permuting the six variables (and the number above the sum is the number of terms). This is visibly invariant under the Weyl group <span class=\"math-container\">$D_6$</span> generated by arbitrary permutations and even numbers of sign changes.</p>\n" }, { "answer_id": 259680, "score": 39, "text": "<p>These are the so-called Regge symmetries, described by T. Regge in a 1970-ish paper. For a bit on it, with references, see the paper</p>\n\n<p><em>Philip P. Boalch</em>, MR 2342290 <a href=\"http://dx.doi.org/10.1007/s00220-007-0328-x\"><strong>Regge and Okamoto symmetries</strong></a>, <em>Comm. Math. Phys.</em> <strong>276</strong> (2007), no. 1, 117--130.\\</p>\n" }, { "answer_id": 259767, "score": 21, "text": "<p>To address the scissors congruence question at the end of the post: the Regge symmetries produce tetrahedra which are scissors congruent. This is proved in Section 6 (Theorem 9 and Corollary 10) of </p>\n\n<ul>\n<li>J. Roberts. Classical <span class=\"math-container\">$6j$</span>-symbols and the tetrahedron. Geom. &amp; Top. 3 (1999), pp. 21-66. <a href=\"http://arxiv.org/abs/math-ph/9812013\" rel=\"nofollow noreferrer\">(link to paper on arXiv)</a></li>\n</ul>\n\n<p>The argument is indirect, proving that the Regge symmetries preserve volume and Dehn invariant, then deducing the statement on scissors congruence from Sydler's theorem.</p>\n\n<p>The corresponding statement for hyperbolic tetrahedra was proved (by construction) in </p>\n\n<ul>\n<li>Y. Mohanty. The Regge symmetry is a scissors congruence in hyperbolic space. Alg. Geom. Top. 3 (2003), 1-31. <a href=\"https://arxiv.org/abs/math/0301318\" rel=\"nofollow noreferrer\">(link to paper on arXiv)</a></li>\n</ul>\n" }, { "answer_id": 259989, "score": 19, "text": "<p>Let me give a geometric interpretation for the case of tetrahedra of volume zero. The statement becomes as follows:</p>\n\n<p>Given four positive numbers $a,b,c,d$ that satisfy the quadrangle inequalities, denote\n$$\ns=\\frac{a+b+c+d}2, \\quad a'=s-a \\text{ etc.}\n$$\nTake a quadrilateral with side lengths $a,b,c,d$ (in this cyclic order). If its diagonals have lengths $x,y$, then there exists a quadrilateral with side lengths $a',b',c',d'$ and the same diagonal lengths.</p>\n\n<p>Again, this can be proved by a direct computation, but Arseniy Akopyan noticed that this is equivalent to the Ivory lemma: diagonals in a curvilinear quadrilateral formed by four confocal conics are equal, see the left half of the picture.</p>\n\n<p><a href=\"https://i.stack.imgur.com/7KCQP.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/7KCQP.png\" alt=\"enter image description here\"></a></p>\n\n<p>Indeed, with two opposite vertices of an $(a,b,c,d)$-quadrilateral as the foci, draw two ellipses and two hyperbolas through the other two vertices. This gives a curvilinear quadrilateral whose diagonal is also a diagonal of our $(a,b,c,d)$-quadrilateral, see the right half of the picture. The relations between $a,b,c,d$ and $a',b',c',d'$:\n$$\na+b=c'+d', c-d=d'-c' \\text{ etc.}\n$$\nimply that the other two vertices of this curvilinear quadrilateral form together with the foci an $(a',b',c',d')$-quadrilateral.</p>\n\n<p>As a side remark: I studied configuration spaces of planar quadrilaterals in <a href=\"https://arxiv.org/abs/1501.07157\" rel=\"noreferrer\" title=\"this preprint\">this preprint</a>, but didn't know about Regge symmetries. Thanks to Dylan and Igor for bringing this up!</p>\n\n<p>There is an analog of the Ivory lemma in dimension three, but there is no word about volumes in it. Still, it would be interesting to look at the corresponding configuration of confocal surfaces (one of the families degenerates).</p>\n\n<p>Also note that the Ivory lemma holds in the hyperbolic and spherical geometry as well.</p>\n" }, { "answer_id": 259847, "score": 7, "text": "<p>It may be worth noting that for some choices of six lengths that there exist as many as 30 inequivalent tetrahedra with these six lengths. </p>\n" }, { "answer_id": 376730, "score": 4, "text": "<p>The following comment in the question intrigued me:</p>\n<blockquote>\n<p>In fact, it's possible to show that the linear symmetries of\n<span class=\"math-container\">$\\mathbb{R}^6$</span> that preserve the Cayley-Menger determinant form the\nWeyl group <span class=\"math-container\">$D_6$</span>, of order <span class=\"math-container\">$2^5 * 6! = 23040$</span>.</p>\n</blockquote>\n<p>In particular, this suggests that it should be possible to perform a change of basis such that the Cayley–Menger determinant is a symmetric polynomial in the six variables, which is also invariant under even numbers of sign changes.</p>\n<p>Some experimentation confirmed that the following variables work in the sense that the geometric and Regge symmetries all preserve the 12-element set <span class=\"math-container\">$\\{ \\pm u, \\pm v, \\pm w, \\pm x, \\pm y, \\pm z \\}$</span>:</p>\n<ul>\n<li><span class=\"math-container\">$u = \\ell_{01} + \\ell_{23}$</span>;</li>\n<li><span class=\"math-container\">$v = \\ell_{02} + \\ell_{13}$</span>;</li>\n<li><span class=\"math-container\">$w = \\ell_{03} + \\ell_{12}$</span>;</li>\n<li><span class=\"math-container\">$x = \\ell_{01} - \\ell_{23}$</span>;</li>\n<li><span class=\"math-container\">$y = \\ell_{02} - \\ell_{13}$</span>;</li>\n<li><span class=\"math-container\">$z = \\ell_{03} - \\ell_{12}$</span>;</li>\n</ul>\n<p>More succinctly, the variables are the three sums of opposite edges and the three differences of opposite edges. The Cayley–Menger determinant can then be shown to be equal to the following 57-term polynomial:</p>\n<p><span class=\"math-container\">$$ uvwxyz - \\dfrac{1}{16} \\sum_\\text{sym}^{6} u^6 + \\dfrac{1}{16} \\sum_\\text{sym}^{30} u^4 v^2 - \\dfrac{1}{8} \\sum_\\text{sym}^{20} u^2 v^2 w^2 $$</span></p>\n<p>where the ‘symmetric sums’ are over all distinct monomials obtained by permuting the six variables (and the number above the sum is the number of terms). This is visibly invariant under the Weyl group <span class=\"math-container\">$D_6$</span> generated by arbitrary permutations and even numbers of sign changes.</p>\n" } ]
Question: Unusual symmetries of the Cayley-Menger determinant for the volume of tetrahedra Answer 1: <p>These are the so-called Regge symmetries, described by T. Regge in a 1970-ish paper. For a bit on it, with references, see the paper</p> <p><em>Philip P. Boalch</em>, MR 2342290 <a href="http://dx.doi.org/10.1007/s00220-007-0328-x"><strong>Regge and Okamoto symmetries</strong></a>, <em>Comm. Math. Phys.</em> <strong>276</strong> (2007), no. 1, 117--130.\</p> Answer 2: <p>To address the scissors congruence question at the end of the post: the Regge symmetries produce tetrahedra which are scissors congruent. This is proved in Section 6 (Theorem 9 and Corollary 10) of </p> <ul> <li>J. Roberts. Classical <span class="math-container">$6j$</span>-symbols and the tetrahedron. Geom. &amp; Top. 3 (1999), pp. 21-66. <a href="http://arxiv.org/abs/math-ph/9812013" rel="nofollow noreferrer">(link to paper on arXiv)</a></li> </ul> <p>The argument is indirect, proving that the Regge symmetries preserve volume and Dehn invariant, then deducing the statement on scissors congruence from Sydler's theorem.</p> <p>The corresponding statement for hyperbolic tetrahedra was proved (by construction) in </p> <ul> <li>Y. Mohanty. The Regge symmetry is a scissors congruence in hyperbolic space. Alg. Geom. Top. 3 (2003), 1-31. <a href="https://arxiv.org/abs/math/0301318" rel="nofollow noreferrer">(link to paper on arXiv)</a></li> </ul> Answer 3: <p>Let me give a geometric interpretation for the case of tetrahedra of volume zero. The statement becomes as follows:</p> <p>Given four positive numbers $a,b,c,d$ that satisfy the quadrangle inequalities, denote $$ s=\frac{a+b+c+d}2, \quad a'=s-a \text{ etc.} $$ Take a quadrilateral with side lengths $a,b,c,d$ (in this cyclic order). If its diagonals have lengths $x,y$, then there exists a quadrilateral with side lengths $a',b',c',d'$ and the same diagonal lengths.</p> <p>Again, this can be proved by a direct computation, but Arseniy Akopyan noticed that this is equivalent to the Ivory lemma: diagonals in a curvilinear quadrilateral formed by four confocal conics are equal, see the left half of the picture.</p> <p><a href="https://i.stack.imgur.com/7KCQP.png" rel="noreferrer"><img src="https://i.stack.imgur.com/7KCQP.png" alt="enter image description here"></a></p> <p>Indeed, with two opposite vertices of an $(a,b,c,d)$-quadrilateral as the foci, draw two ellipses and two hyperbolas through the other two vertices. This gives a curvilinear quadrilateral whose diagonal is also a diagonal of our $(a,b,c,d)$-quadrilateral, see the right half of the picture. The relations between $a,b,c,d$ and $a',b',c',d'$: $$ a+b=c'+d', c-d=d'-c' \text{ etc.} $$ imply that the other two vertices of this curvilinear quadrilateral form together with the foci an $(a',b',c',d')$-quadrilateral.</p> <p>As a side remark: I studied configuration spaces of planar quadrilaterals in <a href="https://arxiv.org/abs/1501.07157" rel="noreferrer" title="this preprint">this preprint</a>, but didn't know about Regge symmetries. Thanks to Dylan and Igor for bringing this up!</p> <p>There is an analog of the Ivory lemma in dimension three, but there is no word about volumes in it. Still, it would be interesting to look at the corresponding configuration of confocal surfaces (one of the families degenerates).</p> <p>Also note that the Ivory lemma holds in the hyperbolic and spherical geometry as well.</p> Answer 4: <p>It may be worth noting that for some choices of six lengths that there exist as many as 30 inequivalent tetrahedra with these six lengths. </p> Answer 5: <p>The following comment in the question intrigued me:</p> <blockquote> <p>In fact, it's possible to show that the linear symmetries of <span class="math-container">$\mathbb{R}^6$</span> that preserve the Cayley-Menger determinant form the Weyl group <span class="math-container">$D_6$</span>, of order <span class="math-container">$2^5 * 6! = 23040$</span>.</p> </blockquote> <p>In particular, this suggests that it should be possible to perform a change of basis such that the Cayley–Menger determinant is a symmetric polynomial in the six variables, which is also invariant under even numbers of sign changes.</p> <p>Some experimentation confirmed that the following variables work in the sense that the geometric and Regge symmetries all preserve the 12-element set <span class="math-container">$\{ \pm u, \pm v, \pm w, \pm x, \pm y, \pm z \}$</span>:</p> <ul> <li><span class="math-container">$u = \ell_{01} + \ell_{23}$</span>;</li> <li><span class="math-container">$v = \ell_{02} + \ell_{13}$</span>;</li> <li><span class="math-container">$w = \ell_{03} + \ell_{12}$</span>;</li> <li><span class="math-container">$x = \ell_{01} - \ell_{23}$</span>;</li> <li><span class="math-container">$y = \ell_{02} - \ell_{13}$</span>;</li> <li><span class="math-container">$z = \ell_{03} - \ell_{12}$</span>;</li> </ul> <p>More succinctly, the variables are the three sums of opposite edges and the three differences of opposite edges. The Cayley–Menger determinant can then be shown to be equal to the following 57-term polynomial:</p> <p><span class="math-container">$$ uvwxyz - \dfrac{1}{16} \sum_\text{sym}^{6} u^6 + \dfrac{1}{16} \sum_\text{sym}^{30} u^4 v^2 - \dfrac{1}{8} \sum_\text{sym}^{20} u^2 v^2 w^2 $$</span></p> <p>where the ‘symmetric sums’ are over all distinct monomials obtained by permuting the six variables (and the number above the sum is the number of terms). This is visibly invariant under the Weyl group <span class="math-container">$D_6$</span> generated by arbitrary permutations and even numbers of sign changes.</p> Answer 6: <p>These are the so-called Regge symmetries, described by T. Regge in a 1970-ish paper. For a bit on it, with references, see the paper</p> <p><em>Philip P. Boalch</em>, MR 2342290 <a href="http://dx.doi.org/10.1007/s00220-007-0328-x"><strong>Regge and Okamoto symmetries</strong></a>, <em>Comm. Math. Phys.</em> <strong>276</strong> (2007), no. 1, 117--130.\</p> Answer 7: <p>To address the scissors congruence question at the end of the post: the Regge symmetries produce tetrahedra which are scissors congruent. This is proved in Section 6 (Theorem 9 and Corollary 10) of </p> <ul> <li>J. Roberts. Classical <span class="math-container">$6j$</span>-symbols and the tetrahedron. Geom. &amp; Top. 3 (1999), pp. 21-66. <a href="http://arxiv.org/abs/math-ph/9812013" rel="nofollow noreferrer">(link to paper on arXiv)</a></li> </ul> <p>The argument is indirect, proving that the Regge symmetries preserve volume and Dehn invariant, then deducing the statement on scissors congruence from Sydler's theorem.</p> <p>The corresponding statement for hyperbolic tetrahedra was proved (by construction) in </p> <ul> <li>Y. Mohanty. The Regge symmetry is a scissors congruence in hyperbolic space. Alg. Geom. Top. 3 (2003), 1-31. <a href="https://arxiv.org/abs/math/0301318" rel="nofollow noreferrer">(link to paper on arXiv)</a></li> </ul> Answer 8: <p>Let me give a geometric interpretation for the case of tetrahedra of volume zero. The statement becomes as follows:</p> <p>Given four positive numbers $a,b,c,d$ that satisfy the quadrangle inequalities, denote $$ s=\frac{a+b+c+d}2, \quad a'=s-a \text{ etc.} $$ Take a quadrilateral with side lengths $a,b,c,d$ (in this cyclic order). If its diagonals have lengths $x,y$, then there exists a quadrilateral with side lengths $a',b',c',d'$ and the same diagonal lengths.</p> <p>Again, this can be proved by a direct computation, but Arseniy Akopyan noticed that this is equivalent to the Ivory lemma: diagonals in a curvilinear quadrilateral formed by four confocal conics are equal, see the left half of the picture.</p> <p><a href="https://i.stack.imgur.com/7KCQP.png" rel="noreferrer"><img src="https://i.stack.imgur.com/7KCQP.png" alt="enter image description here"></a></p> <p>Indeed, with two opposite vertices of an $(a,b,c,d)$-quadrilateral as the foci, draw two ellipses and two hyperbolas through the other two vertices. This gives a curvilinear quadrilateral whose diagonal is also a diagonal of our $(a,b,c,d)$-quadrilateral, see the right half of the picture. The relations between $a,b,c,d$ and $a',b',c',d'$: $$ a+b=c'+d', c-d=d'-c' \text{ etc.} $$ imply that the other two vertices of this curvilinear quadrilateral form together with the foci an $(a',b',c',d')$-quadrilateral.</p> <p>As a side remark: I studied configuration spaces of planar quadrilaterals in <a href="https://arxiv.org/abs/1501.07157" rel="noreferrer" title="this preprint">this preprint</a>, but didn't know about Regge symmetries. Thanks to Dylan and Igor for bringing this up!</p> <p>There is an analog of the Ivory lemma in dimension three, but there is no word about volumes in it. Still, it would be interesting to look at the corresponding configuration of confocal surfaces (one of the families degenerates).</p> <p>Also note that the Ivory lemma holds in the hyperbolic and spherical geometry as well.</p> Answer 9: <p>It may be worth noting that for some choices of six lengths that there exist as many as 30 inequivalent tetrahedra with these six lengths. </p> Answer 10: <p>The following comment in the question intrigued me:</p> <blockquote> <p>In fact, it's possible to show that the linear symmetries of <span class="math-container">$\mathbb{R}^6$</span> that preserve the Cayley-Menger determinant form the Weyl group <span class="math-container">$D_6$</span>, of order <span class="math-container">$2^5 * 6! = 23040$</span>.</p> </blockquote> <p>In particular, this suggests that it should be possible to perform a change of basis such that the Cayley–Menger determinant is a symmetric polynomial in the six variables, which is also invariant under even numbers of sign changes.</p> <p>Some experimentation confirmed that the following variables work in the sense that the geometric and Regge symmetries all preserve the 12-element set <span class="math-container">$\{ \pm u, \pm v, \pm w, \pm x, \pm y, \pm z \}$</span>:</p> <ul> <li><span class="math-container">$u = \ell_{01} + \ell_{23}$</span>;</li> <li><span class="math-container">$v = \ell_{02} + \ell_{13}$</span>;</li> <li><span class="math-container">$w = \ell_{03} + \ell_{12}$</span>;</li> <li><span class="math-container">$x = \ell_{01} - \ell_{23}$</span>;</li> <li><span class="math-container">$y = \ell_{02} - \ell_{13}$</span>;</li> <li><span class="math-container">$z = \ell_{03} - \ell_{12}$</span>;</li> </ul> <p>More succinctly, the variables are the three sums of opposite edges and the three differences of opposite edges. The Cayley–Menger determinant can then be shown to be equal to the following 57-term polynomial:</p> <p><span class="math-container">$$ uvwxyz - \dfrac{1}{16} \sum_\text{sym}^{6} u^6 + \dfrac{1}{16} \sum_\text{sym}^{30} u^4 v^2 - \dfrac{1}{8} \sum_\text{sym}^{20} u^2 v^2 w^2 $$</span></p> <p>where the ‘symmetric sums’ are over all distinct monomials obtained by permuting the six variables (and the number above the sum is the number of terms). This is visibly invariant under the Weyl group <span class="math-container">$D_6$</span> generated by arbitrary permutations and even numbers of sign changes.</p> Answer 11: <p>These are the so-called Regge symmetries, described by T. Regge in a 1970-ish paper. For a bit on it, with references, see the paper</p> <p><em>Philip P. Boalch</em>, MR 2342290 <a href="http://dx.doi.org/10.1007/s00220-007-0328-x"><strong>Regge and Okamoto symmetries</strong></a>, <em>Comm. Math. Phys.</em> <strong>276</strong> (2007), no. 1, 117--130.\</p> Answer 12: <p>To address the scissors congruence question at the end of the post: the Regge symmetries produce tetrahedra which are scissors congruent. This is proved in Section 6 (Theorem 9 and Corollary 10) of </p> <ul> <li>J. Roberts. Classical <span class="math-container">$6j$</span>-symbols and the tetrahedron. Geom. &amp; Top. 3 (1999), pp. 21-66. <a href="http://arxiv.org/abs/math-ph/9812013" rel="nofollow noreferrer">(link to paper on arXiv)</a></li> </ul> <p>The argument is indirect, proving that the Regge symmetries preserve volume and Dehn invariant, then deducing the statement on scissors congruence from Sydler's theorem.</p> <p>The corresponding statement for hyperbolic tetrahedra was proved (by construction) in </p> <ul> <li>Y. Mohanty. The Regge symmetry is a scissors congruence in hyperbolic space. Alg. Geom. Top. 3 (2003), 1-31. <a href="https://arxiv.org/abs/math/0301318" rel="nofollow noreferrer">(link to paper on arXiv)</a></li> </ul> Answer 13: <p>Let me give a geometric interpretation for the case of tetrahedra of volume zero. The statement becomes as follows:</p> <p>Given four positive numbers $a,b,c,d$ that satisfy the quadrangle inequalities, denote $$ s=\frac{a+b+c+d}2, \quad a'=s-a \text{ etc.} $$ Take a quadrilateral with side lengths $a,b,c,d$ (in this cyclic order). If its diagonals have lengths $x,y$, then there exists a quadrilateral with side lengths $a',b',c',d'$ and the same diagonal lengths.</p> <p>Again, this can be proved by a direct computation, but Arseniy Akopyan noticed that this is equivalent to the Ivory lemma: diagonals in a curvilinear quadrilateral formed by four confocal conics are equal, see the left half of the picture.</p> <p><a href="https://i.stack.imgur.com/7KCQP.png" rel="noreferrer"><img src="https://i.stack.imgur.com/7KCQP.png" alt="enter image description here"></a></p> <p>Indeed, with two opposite vertices of an $(a,b,c,d)$-quadrilateral as the foci, draw two ellipses and two hyperbolas through the other two vertices. This gives a curvilinear quadrilateral whose diagonal is also a diagonal of our $(a,b,c,d)$-quadrilateral, see the right half of the picture. The relations between $a,b,c,d$ and $a',b',c',d'$: $$ a+b=c'+d', c-d=d'-c' \text{ etc.} $$ imply that the other two vertices of this curvilinear quadrilateral form together with the foci an $(a',b',c',d')$-quadrilateral.</p> <p>As a side remark: I studied configuration spaces of planar quadrilaterals in <a href="https://arxiv.org/abs/1501.07157" rel="noreferrer" title="this preprint">this preprint</a>, but didn't know about Regge symmetries. Thanks to Dylan and Igor for bringing this up!</p> <p>There is an analog of the Ivory lemma in dimension three, but there is no word about volumes in it. Still, it would be interesting to look at the corresponding configuration of confocal surfaces (one of the families degenerates).</p> <p>Also note that the Ivory lemma holds in the hyperbolic and spherical geometry as well.</p> Answer 14: <p>It may be worth noting that for some choices of six lengths that there exist as many as 30 inequivalent tetrahedra with these six lengths. </p> Answer 15: <p>The following comment in the question intrigued me:</p> <blockquote> <p>In fact, it's possible to show that the linear symmetries of <span class="math-container">$\mathbb{R}^6$</span> that preserve the Cayley-Menger determinant form the Weyl group <span class="math-container">$D_6$</span>, of order <span class="math-container">$2^5 * 6! = 23040$</span>.</p> </blockquote> <p>In particular, this suggests that it should be possible to perform a change of basis such that the Cayley–Menger determinant is a symmetric polynomial in the six variables, which is also invariant under even numbers of sign changes.</p> <p>Some experimentation confirmed that the following variables work in the sense that the geometric and Regge symmetries all preserve the 12-element set <span class="math-container">$\{ \pm u, \pm v, \pm w, \pm x, \pm y, \pm z \}$</span>:</p> <ul> <li><span class="math-container">$u = \ell_{01} + \ell_{23}$</span>;</li> <li><span class="math-container">$v = \ell_{02} + \ell_{13}$</span>;</li> <li><span class="math-container">$w = \ell_{03} + \ell_{12}$</span>;</li> <li><span class="math-container">$x = \ell_{01} - \ell_{23}$</span>;</li> <li><span class="math-container">$y = \ell_{02} - \ell_{13}$</span>;</li> <li><span class="math-container">$z = \ell_{03} - \ell_{12}$</span>;</li> </ul> <p>More succinctly, the variables are the three sums of opposite edges and the three differences of opposite edges. The Cayley–Menger determinant can then be shown to be equal to the following 57-term polynomial:</p> <p><span class="math-container">$$ uvwxyz - \dfrac{1}{16} \sum_\text{sym}^{6} u^6 + \dfrac{1}{16} \sum_\text{sym}^{30} u^4 v^2 - \dfrac{1}{8} \sum_\text{sym}^{20} u^2 v^2 w^2 $$</span></p> <p>where the ‘symmetric sums’ are over all distinct monomials obtained by permuting the six variables (and the number above the sum is the number of terms). This is visibly invariant under the Weyl group <span class="math-container">$D_6$</span> generated by arbitrary permutations and even numbers of sign changes.</p> Answer 16: <p>These are the so-called Regge symmetries, described by T. Regge in a 1970-ish paper. For a bit on it, with references, see the paper</p> <p><em>Philip P. Boalch</em>, MR 2342290 <a href="http://dx.doi.org/10.1007/s00220-007-0328-x"><strong>Regge and Okamoto symmetries</strong></a>, <em>Comm. Math. Phys.</em> <strong>276</strong> (2007), no. 1, 117--130.\</p> Answer 17: <p>To address the scissors congruence question at the end of the post: the Regge symmetries produce tetrahedra which are scissors congruent. This is proved in Section 6 (Theorem 9 and Corollary 10) of </p> <ul> <li>J. Roberts. Classical <span class="math-container">$6j$</span>-symbols and the tetrahedron. Geom. &amp; Top. 3 (1999), pp. 21-66. <a href="http://arxiv.org/abs/math-ph/9812013" rel="nofollow noreferrer">(link to paper on arXiv)</a></li> </ul> <p>The argument is indirect, proving that the Regge symmetries preserve volume and Dehn invariant, then deducing the statement on scissors congruence from Sydler's theorem.</p> <p>The corresponding statement for hyperbolic tetrahedra was proved (by construction) in </p> <ul> <li>Y. Mohanty. The Regge symmetry is a scissors congruence in hyperbolic space. Alg. Geom. Top. 3 (2003), 1-31. <a href="https://arxiv.org/abs/math/0301318" rel="nofollow noreferrer">(link to paper on arXiv)</a></li> </ul> Answer 18: <p>Let me give a geometric interpretation for the case of tetrahedra of volume zero. The statement becomes as follows:</p> <p>Given four positive numbers $a,b,c,d$ that satisfy the quadrangle inequalities, denote $$ s=\frac{a+b+c+d}2, \quad a'=s-a \text{ etc.} $$ Take a quadrilateral with side lengths $a,b,c,d$ (in this cyclic order). If its diagonals have lengths $x,y$, then there exists a quadrilateral with side lengths $a',b',c',d'$ and the same diagonal lengths.</p> <p>Again, this can be proved by a direct computation, but Arseniy Akopyan noticed that this is equivalent to the Ivory lemma: diagonals in a curvilinear quadrilateral formed by four confocal conics are equal, see the left half of the picture.</p> <p><a href="https://i.stack.imgur.com/7KCQP.png" rel="noreferrer"><img src="https://i.stack.imgur.com/7KCQP.png" alt="enter image description here"></a></p> <p>Indeed, with two opposite vertices of an $(a,b,c,d)$-quadrilateral as the foci, draw two ellipses and two hyperbolas through the other two vertices. This gives a curvilinear quadrilateral whose diagonal is also a diagonal of our $(a,b,c,d)$-quadrilateral, see the right half of the picture. The relations between $a,b,c,d$ and $a',b',c',d'$: $$ a+b=c'+d', c-d=d'-c' \text{ etc.} $$ imply that the other two vertices of this curvilinear quadrilateral form together with the foci an $(a',b',c',d')$-quadrilateral.</p> <p>As a side remark: I studied configuration spaces of planar quadrilaterals in <a href="https://arxiv.org/abs/1501.07157" rel="noreferrer" title="this preprint">this preprint</a>, but didn't know about Regge symmetries. Thanks to Dylan and Igor for bringing this up!</p> <p>There is an analog of the Ivory lemma in dimension three, but there is no word about volumes in it. Still, it would be interesting to look at the corresponding configuration of confocal surfaces (one of the families degenerates).</p> <p>Also note that the Ivory lemma holds in the hyperbolic and spherical geometry as well.</p> Answer 19: <p>It may be worth noting that for some choices of six lengths that there exist as many as 30 inequivalent tetrahedra with these six lengths. </p> Answer 20: <p>The following comment in the question intrigued me:</p> <blockquote> <p>In fact, it's possible to show that the linear symmetries of <span class="math-container">$\mathbb{R}^6$</span> that preserve the Cayley-Menger determinant form the Weyl group <span class="math-container">$D_6$</span>, of order <span class="math-container">$2^5 * 6! = 23040$</span>.</p> </blockquote> <p>In particular, this suggests that it should be possible to perform a change of basis such that the Cayley–Menger determinant is a symmetric polynomial in the six variables, which is also invariant under even numbers of sign changes.</p> <p>Some experimentation confirmed that the following variables work in the sense that the geometric and Regge symmetries all preserve the 12-element set <span class="math-container">$\{ \pm u, \pm v, \pm w, \pm x, \pm y, \pm z \}$</span>:</p> <ul> <li><span class="math-container">$u = \ell_{01} + \ell_{23}$</span>;</li> <li><span class="math-container">$v = \ell_{02} + \ell_{13}$</span>;</li> <li><span class="math-container">$w = \ell_{03} + \ell_{12}$</span>;</li> <li><span class="math-container">$x = \ell_{01} - \ell_{23}$</span>;</li> <li><span class="math-container">$y = \ell_{02} - \ell_{13}$</span>;</li> <li><span class="math-container">$z = \ell_{03} - \ell_{12}$</span>;</li> </ul> <p>More succinctly, the variables are the three sums of opposite edges and the three differences of opposite edges. The Cayley–Menger determinant can then be shown to be equal to the following 57-term polynomial:</p> <p><span class="math-container">$$ uvwxyz - \dfrac{1}{16} \sum_\text{sym}^{6} u^6 + \dfrac{1}{16} \sum_\text{sym}^{30} u^4 v^2 - \dfrac{1}{8} \sum_\text{sym}^{20} u^2 v^2 w^2 $$</span></p> <p>where the ‘symmetric sums’ are over all distinct monomials obtained by permuting the six variables (and the number above the sum is the number of terms). This is visibly invariant under the Weyl group <span class="math-container">$D_6$</span> generated by arbitrary permutations and even numbers of sign changes.</p>
The missing link: an inequality
259,197
44
https://mathoverflow.net/questions/259197/the-missing-link-an-inequality
<p>I've been working on a project and proved a few relevant results, but got stuck on one tricky problem:</p> <blockquote> <p><strong>Conjecture.</strong> If $2\leq n\in\mathbb{N}$ and $0&lt;x&lt;1$ is a real number, then $$F_n(x)=\log\left(\frac{(1+x^{4n-1})(1+x^{2n})(1-x^{2n+1})}{(1+x^{2n+1})(1-x^{2n+2})}\right)$$ is a convex function of $x$.</p> <p>I have a heuristic argument. Can you help with a rigorous proof or valuable tools? </p> </blockquote> <p><strong>Further motivation.</strong> If you succeed with this, then I'll be honored to have you as a <a href="https://www.math.temple.edu/~tewodros/Catalan_Convexity.pdf">co-author in this work</a>. The problem itself can be found in Section 4.</p> <p><strong>Note.</strong> Write $F=\log\frac{P}Q$, then $F''&gt;0$ amounts to the positivity of the polynomial $$V:=PQ^2P''+(PQ')^2-P^2QQ''-(P'Q)^2.$$</p>
[ { "answer_id": 259370, "score": 21, "text": "<p>Fix $n$, and let $V(x)$ be the polynomial from the question, for which we want to show that it is\npositive on the open interval $(0,1)$. Set\n\\begin{align*}\na(n) &amp;= 1024n^2 - 4096/3n + 320\\\\\nb(n) &amp;= 12288n^3 - 14336n^2 + 4864/3n + 256\\\\\nW(x) &amp;= x^4\\left(V(x)-(n+1)^2(2n+1)^2x^{24n-1}(1-x)^4(a(n)x + b(n)(1-x))\\right).\n\\end{align*}</p>\n\n<p>For all $n\\ge12$ we claim that $(1-x)^6$ divides $W(x)$, and that all the coefficients of $W(x)/(1-x)^6$ are nonnegative. In particular, $W(x)$ is positive for $x&gt;0$. As $a(n)$ and $b(n)$ is positive for all $n\\ge1$, we see that $V(x)$ is positive for all $0&lt;x&lt;1$, provided that $n\\ge12$. The cases $2\\le n\\le 11$ are easily checked directly, for instance by noting that the coefficients of $(1+x)^{\\deg V}V(1/(1+x))$ are positive.</p>\n\n<p>I'm not sure about the easiest kind to prove the assertion about the coefficients of $W(x)/(1-x)^6$. We compute them via\n\\begin{equation}\n\\frac{W(x)}{(1-x)^6}=W(x)\\sum_{k\\ge0}\\binom{k+5}{5}x^k.\n\\end{equation}</p>\n\n<p>All what follows is assisted/confirmed by the Sage code below.</p>\n\n<p>The exponents of $W(x)$ have the form $2ni+j$ for $0\\le i\\le 12$ and $0\\le j\\le 10$. Let $a_{i,j}$ be the corresponding coefficient. It is a polynomial in $n$.</p>\n\n<p>We have\n\\begin{equation} \n\\frac{W(x)}{(1-x)^6} = \\sum_{i,j}a_{i,j}(n)x^{2ni+j}\\sum_k\\binom{k+5}{5}x^k.\n\\end{equation}</p>\n\n<p>With $r\\ge0$ and $0\\le s\\le 2n-1$, the coefficient of $x^{2nr+s}$ in $W(x)\\frac{1}{(1-x)^6}$\nis the sum of all $a_{i,j}(n)\\cdot\\binom{nr+s-ni-j+y}{5}$ for all pairs $(i,j)$ where $i&lt;r$ and $0\\le j\\le 10$, or $i=r$ and $j\\le s$.</p>\n\n<p>We distinguish the cases $0\\le s\\le 9$ from the cases $s\\ge10$. In these latter cases, we write $s=10+y$. Recall that $s\\le 2n-1$, so $10+y=2n-1-o$ for a nonnegative $o$. Upon replacing $n$ with $(11+y+1)/2$, we get polynomials in $n$ and $o$.</p>\n\n<p>In the former cases, the coefficients are polynomials in $n$, which have positive coefficients upon replacing $n$ with $n+12$. So these coefficients are positive for all $n\\ge12$. </p>\n\n<p>In the latter cases it turns out that in all but one case the coefficients are positive. So for nonnegative $y$ and $o$, the values are nonnegative.</p>\n\n<p>For the single exception we see that if we multiply it with a suitable polynomial, the resulting coefficients are positive.</p>\n\n<p><em>Remark (answering Jason's question):</em> It is a known fact that a polynomial $V$ is positive on $(0,1)$ if and only if it is a nonnegative linear combination of polynomials $x^i(1-x)^j$. The problem is that it may involve terms where $i+j$ is bigger than $\\deg V$. (This doesn't happen here, though.) I used an LP solver to play a little with $V$ and $\\frac{V}{(1-x^2)^4x^{2n-2}}$. From that a pattern showed up which lead to a solution.</p>\n\n<p>By the way, $V$ actually seems to be a nonnegative linear combination of $x^i(1-x)^{\\deg V-i}$. This is equivalent to say that all the coefficients of $(1+x)^{\\deg V}V(1/(1+x))$ are nonnegative. It is not hard to compute explicit expressions for these coefficients, but I don't see an easy arguments why they can't be negative.</p>\n\n<pre><code># Formally compute W\nvar('X N')\nP = (1+X^(4*N-1))*(1+X^(2*N))*(1-X^(2*N+1))\nQ = (1+X^(2*N+1))*(1-X^(2*N+2))\nV = P*Q^2*P.diff(X,2)+(P*Q.diff(X))^2-P^2*Q*Q.diff(X,2)-(P.diff(X)*Q)^2\na = 1024*N^2 - 4096/3*N + 320\nb = 12288*N^3 - 14336*N^2 + 4864/3*N + 256\nW = X^4*(V - (2*N + 1)^2 *(N + 1)^2*X^(24*N-1)*(1-X)^4*(a*X + b*(1-X)))\n\n# Check that (1-X)^6 divides W(X)\nprint all(W.diff(X,i)(X=1).polynomial(QQ) == 0 for i in [0..5])\n\n# Compute the coefficients a_ij for the exponents 2ni+j of W. Somewhat\n# clumsy, as I don't know how to deal with polynomials where exponents\n# are symbolic expressions.\n#\n# l is the list of summands of W\nl = [z.canonicalize_radical() for z in W.expand().operands()]\nK.&lt;n&gt; = QQ[]\naij = {(i,j):K(0) for i in [0..12] for j in [0..10]}\nfor term in l:\n c = term(X=1) # get coefficient of term\n e = (term.diff(X)/c)(X=1) # get exponent of term\n c = K(c.polynomial(QQ)) # convert c to proper polynomial in n\n i, j = ZZ(e.diff(N))//2, ZZ(e(N=0)) # Clumsy method to compute pairs (i,j)\n aij[i,j] += c\n\n# Check if coefficients aij[i,j] were correctly computed\nWnew = sum(c(n=N)*X^(2*N*i+j) for (i,j),c in aij.items())\nprint (W-Wnew).canonicalize_radical().is_trivial_zero()\n\ndef bino(k): # binomial coefficient binom(-6,k)=binom(k+5,5)\n return prod(k+5-z for z in range(5))/120\n\n# compute the coefficients of W(X)/(1-X)^6\nK = K.extend_variables(('y', 'o'))\nK.inject_variables()\nfor r in [0..12]:\n for s in [0..10]:\n d = 1 if s == 10 else 0\n # compute coefficient of X^(2nr+s+d*y)\n f = sum(aij[i,j]*bino(2*n*(r-i)+s+d*y-j) for i in [0..r]\n for j in [0..s if i == r else 10])\n # check non-negativity of the polynomial f\n if d == 0:\n # Checks if the coefficient of X^(2nr+s), which is a\n # polynomial in n, has nonnegative coefficients upon replacing\n # n with n+12\n f = f(n=n+12)\n if min(f.coefficients()+[0]) &lt; 0:\n print \"False\"\n else:\n # The coefficient of X^(2nr+10+y), which is a polynomial in n\n # and y, has to be nonnegative for 0 &lt;= 10+y &lt;= 2n-1, so 10+y\n # = 2n-1-o for non-negative o. So upon replacing n with\n # (11+y+o)/2, we get a polynomial in y and o which has to be\n # nonnegative for all nonnegative y and o. In all but one\n # case, this holds because the coeffcients are non-negative.\n f = f(n=(11+y+o)/2)\n if min(f.coefficients()+[0]) &lt; 0:\n c = o^2 + 23*o*y + 1360*y^2 + 99*o + 340*y + 1675\n print min(c.coefficients()+(c*f).coefficients()) &gt;= 0\n</code></pre>\n" }, { "answer_id": 259280, "score": 11, "text": "<p>This proof is mainly based on some of the suggestions in the early version of this answer. Also, we are building here on the comment by Peter Mueller concerning the polynomial $a$ defined below. It turns out that similar techniques can be used for all other nontrivial polynomials arising in the proof, which allowed us to replace multiple uses of Mathematica command Reduce[] in the last version of this answer by showing that the coefficients of the mentioned relevant polynomials are all nonnegative. I have still retained some use of Reduce[] -- but only to quickly make the routine check for $n\\le11$. </p>\n\n<p>Note that \n\\begin{equation}\nG:=\nF_n''(x)\tx^{2-2 n} \\left(x^{2 n}+1\\right)^2 \\left(x^{4 n}+x\\right)^2 \\left(x^{2 n+1}-1\\right)^2\n \\left(x^{2 n+1}+1\\right)^2 \\left(x^{2 n+2}-1\\right)^2\n\\end{equation}\nis a polynomial in $n$, $x$, and $y:=x^{2n}$, of degree $2$ in $n$:\n\\begin{equation}\n\tG=a+bn+cn^2,\n\\end{equation}\nwhere $a,b,c$ are certain polynomials in $x,y$. </p>\n\n<p>The check of $G\\ge0$ for $n=2,\\dots,11$ is straightforward. So, assume $n\\ge12$, whence $0&lt;y&lt;x^{24}&lt;x^{12}$. \nWe also always assume that $0&lt;x&lt;1$, $0&lt;y&lt;x^{24}$ , \n\\begin{equation}\n\tn=\\frac{\\ln y}{2\\ln x}; \\tag{1}\n\\end{equation}\nthe latter relation is of course just another form of the equality $y=x^{2n}$. </p>\n\n<p>Note that \n$0&lt;y&lt;1\\ \\&amp;\\ 0&lt;x&lt;1$ implies $x=1/(1+u)\\ \\&amp;\\ y=1/(1+v)$ for some $u,v\\ge0$. \nAs suggested by Peter Mueller, it turns out that \n\\begin{equation}\n\ta\\; (1 + u)^{10} (1 + v)^{11}|_{x=1/(1 + u), y=1/(1 + v)}\n\\end{equation}\nis a polynomial in $u,v$ with all coefficients nonnegative. \nDetails of all calculations can be seen in the <a href=\"https://www.dropbox.com/s/g87a59uzdsb75m5/SolutionM1.nb?dl=0\" rel=\"nofollow noreferrer\">Mathematica notebook</a> \nand its <a href=\"https://www.dropbox.com/s/k4vlx4brvok099k/SolutionM1.pdf?dl=0\" rel=\"nofollow noreferrer\">pdf image</a></p>\n\n<p>So, \n\\begin{equation}\n\ta\\ge0\n\\end{equation}\nfor all $x,y$ in $(0,1)$. </p>\n\n<p>Next, introduce \n\\begin{equation}\nx_* := \\frac{869}{1000},\\quad \nx_0:=\\frac{7655}{10000},\\quad y_0:= x_0^{24}\\approx0.00164,\\quad\nx_1 := \\frac{985}{1000},\\quad y_1:= x_1^{24}\\approx0.696. \n\\end{equation} </p>\n\n<p>The following cases/subcases/subsubcases are exhaustive: </p>\n\n<p><strong>Case 1: $y\\le y_1$.</strong> Then the condition $0&lt;y\\le y_1\\ \\&amp;\\ 0&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;x^{24}$ implies $x=x_1/(1+u)\\ \\&amp;\\ y=x^{24}/(1+v)$ or $x=x_1+(1-x_1)/(1+u)\\ \\&amp;\\ y=y_1/(1+v)$ for some $u\\ge0$ and $v\\ge0$, depending on whether $x&lt;x_1$ or $x\\ge x_1$. \nIt turns out that \n\\begin{equation}\n\tc\\; (1 + u)^{248} (1 + v)^{10}|_{x=x_1/(1+u)\\ \\&amp;\\ y=x^{24}/(1+v)}\n\\end{equation}\nand \n\\begin{equation}\n\tc\\; (1 + u)^{10} (1 + v)^{10}|_{x=x_1+(1-x_1)/(1+u)\\ \\&amp;\\ y=y_1/(1+v)}\n\\end{equation}\nare each a polynomial in $u,v$ with all coefficients nonnegative. \nSo, \n\\begin{equation}\n\tc\\ge0 \\quad\\text{in Case 1}. \n\\end{equation}\nSo, for the derivative $G'_n$ of $G$ in $n$, we have \n\\begin{equation}\\tag{2}\nG'_n=b+2cn\\ge b+24c=(G'_n)|_{n=12}.\t\n\\end{equation}</p>\n\n<p>Now we have to distinguish two subcases of Case 1, with further subsubcases of Subcase 1.1: </p>\n\n<p><strong>Subcase 1.1: $y\\le y_1$ and ($y\\ge y_0$ or $x\\le x_*$).</strong> </p>\n\n<p><strong>Subsubcase 1.1.1: $y\\le y_1$ and $y\\ge y_0$ and $x\\ge x_1$.</strong> Then $y\\le y_1\\ \\&amp;\\ y\\ge y_0\\ \\&amp;\\ x\\ge x_1 \\ \\&amp;\\ 0&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;x^{24}$ implies \n$$x=x_{111}:=x_1+(1-x_1)/(1+u)\\ \\&amp;\\ y=y_{111}:=y_0+(y_1-y_0)/(1+v)$$\nfor some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=12}\\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=12}\\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p>\n\n<p><strong>Subsubcase 1.1.2: $y\\le y_1$ and $y\\ge y_0$ and $x&lt;x_1$.</strong> Then $x=x_{112}:=x_0+(x_1-x_0)/(1+u)\\ \\&amp;\\ y=y_{112}:=y_0+(x^{24}-y_0)/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112}, y=y_{112}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112},y=y_{112}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p>\n\n<p><strong>Subsubcase 1.1.3: $y\\le y_1$ and $x\\le x_*$.</strong> Then $x=x_{113}:=x_*/(1+u)\\ \\&amp;\\ y=y_{113}:=x^{24}/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p>\n\n<p>So, $G|_{n=12}\\ge0$ and $(G'_n)|_{n=12}\\ge0$ in all the three subsubcases of Subcase 1.1. In view of $(2)$, we have \n$G\\ge0$ in Subcase 1.1. </p>\n\n<p><strong>Subcase 1.2: $y\\le y_1$ and $y&lt;y_0$ and $x&gt;x_*$.</strong> Then, of course, the condition $y\\le y_1$ is redundant. Let here \n\\begin{equation}\n\t\\rho(x):=\\frac7{2(1-x)}. \n\\end{equation}\nThe condition \n$x_*&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;y_0$ implies $x=x_{12}:=x_*+(1-x_*)/(1+u)$ and $y=y_{12}:=y_0/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=\\rho(x)/2}\\; u\\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=\\rho(x)/2}\\; u^2\\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. \nSo, similarly to Subcase 1.1, for all $n\\ge\\rho(x)/2$ we get $G\\ge G|_{n=\\rho(x)/2}\\ge0$. \nIn Subcase 1.2, it remains to note that for our particular $n$, as in $(1)$, we indeed have $n\\ge\\rho(x)/2$. This follows because for $y&lt;y_0$ and $x&gt;x_*$ \n\\begin{equation}\n\t\\ln\\frac1y&gt;\\ln\\frac1{y_0}&gt;1.08\\frac7{2(1-x_*)}\\,\\ln\\frac1{x_*}\n\t&gt;\\frac7{2(1-x)}\\,\\ln\\frac1{x}=\\rho(x)\\ln\\frac1{x}. \n\\end{equation}\nSo, $G\\ge0$ in Subcase 1.2 as well. </p>\n\n<p>It remains to consider </p>\n\n<p><strong>Case 2: $y&gt;y_1$.</strong> Then condition $y&gt;y_1\\ \\&amp;\\ 0&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;x^{24}[&lt;x^{12}]$ implies $x=x_2:=x_1+(1-x_1)/(1+u),y=y_2:=y_1+(x^{12}-y_1)/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\tb\\;(1 + u)^{140} (1 + v)^{11}|_{x=x_{2},y=y_{2}}\n\\end{equation}\nis also a polynomial in $u,v$ with all coefficients nonnegative. \nHence, $b\\ge0$ in Case 2. \nSo, if $c\\ge0$, then trivially $G\\ge0$ (since $a\\ge0$ always). So, without loss of generality, $c&lt;0$ and hence $G$ is concave in $n$. So, it is enough to bracket the $n$ as in $(1)$ between some simple rational expressions $n_1$ and $n_2$ such that $G|_{n=n_1}\\ge0$ and $G|_{n=n_2}\\ge0$. </p>\n\n<p>In Case 2, $y$ is \"close\" to $1$ and hence, in view of inequality $y&lt;x^{24}$, so is $x$: $x&gt;y_1^{1/12}=x_1=0.985$. So, it is a bit more convenient here to change variables $x, y$ to \"small variables\" $u:=1-x$ and $v:=1-y$. \nThen $0&lt;v&lt;1-y_1\\approx0.304$ and $1-u=x&gt;y^{1/24}=(1-v)^{1/24}&gt;1-v/3$, whence $u&lt;v/3$. Let now $t:=u/v$, so that $0&lt;t&lt;1/3$. \nIt follows, in Case 2, that $t=(1/3)/(1+r)$ and $v=(1-y_1)/(1+s)$ for some $r\\ge0$ and $s\\ge0$. </p>\n\n<p>The mentioned brackets $n_1$ and $n_2$ are respectively $\\ell_1/2$ and $\\ell_2/2$, where<br>\n\\begin{equation}\n\t\\ell_1:=\\frac1t,\\quad\\ell_2:=\\frac{(2 - v) (1 - t v)}{t (1 - v) (2 - t v)}. \n\\end{equation}\nIt is not hard to see (details on this are in the same Mathematica notebook) that then indeed $n_1 &lt; n &lt; n_2$. \nIt remains to show that $G|_{n=n_1}\\ge0$ and $G|_{n=n_2}\\ge0$. </p>\n\n<p>It turns out that \n\\begin{equation}\n\t(G'_n)|_{n=\\ell_1/2;\\,x=1-tv,\\,y=1-v;\\,t=(1/3)/(1+r),\\,v=(1-y_1)/(1+s)}\\; (1 + r)^{10} (1 + s)^{19}\n\\end{equation}\nand \n\\begin{equation}\n\t(G'_n)|_{n=\\ell_2/2;\\,x=1-tv,\\,y=1-v;\\,t=(1/3)/(1+r),\\,v=(1-y_1)/(1+s)}\\; (1 + r)^{10} (1 + s)^{19}\\; q\n\\end{equation}\nare each a polynomial in $r,s$ with all coefficients nonnegative, \nwhere \n\\begin{multline*}\n\tq:=\n\t(1 + r)^{10} (1 + s)^{19} \\\\\n\t\\times(11673186598630578538556565100133681446610566511878526881 \\\\ \n\t+ \n 16777216000000000000000000000000000000000000000000000000 s)^2 \\\\ \n\\times\\big(31853088866210192846185521700044560482203522170626175627 \\\\\n+ \n 33554432000000000000000000000000000000000000000000000000 (r + s + \n r s)\\big)^2. \n\\end{multline*}</p>\n\n<p>The proof is complete. </p>\n" }, { "answer_id": 259260, "score": 10, "text": "<p>I sketch a method that should show $F_n$ is convex for $n$ sufficiently large.</p>\n\n<p>A Taylor expansion gives that $F_n(x) = x^{2n} (1-x)^2 + O(x^{4n-1})$, and so $$F_n''(x) = 2x^{2n-2}[2n^2(1-x)^2 + x^2 - n(1-x)(1+3x)] + O(x^{4n-3}).$$\nThis should show that $F_n''(x) &gt; 0$ for $n$ large, provided $x &lt; 1-\\frac{C}{n}$ for some fixed $C&gt;0$.</p>\n\n<p>On the other hand, according to Mathematica, $F_n''(1) = 4n^2 - \\frac{16}{3} n + \\frac{5}{4}$, which has its largest root at $\\frac{8 + \\sqrt{19}}{12} \\approx 1.02991$. (By the way, there doesn't seem to be any reason that $n$ should be restricted to be an integer, and this is in line with some plots I made indicating $F_n(x)$ is convex for $n &gt; 1.03$.)\nMoreover, Mathematica gives $$\\lim_{n\\rightarrow \\infty} n^{-2}F_n''(1-\\frac{y}{n}) = \\frac{16 e^{4y}}{(1+e^{4y})^2} &gt; 0.$$</p>\n\n<p>Putting these two pieces together should be able to show $F_n(x)$ is convex on $(0,1)$ for $n$ sufficiently large. Unfortunately, it might be very painful to work out all the error terms explicitly.</p>\n" }, { "answer_id": 259207, "score": 6, "text": "<p>Not an answer, just a plot of $F_n(x)$ for $n=2,\\ldots,10$:\n<hr />\n&nbsp; &nbsp; &nbsp; &nbsp; &nbsp;\n<a href=\"https://i.stack.imgur.com/jzWwY.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/jzWwY.jpg\" alt=\"Fn(x)\"></a>\n<hr />\n&nbsp; &nbsp; &nbsp; &nbsp; &nbsp;\n<a href=\"https://i.stack.imgur.com/c8gFL.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/c8gFL.jpg\" alt=\"Fn(x)2\"></a></p>\n\n<hr />\n" }, { "answer_id": 259213, "score": 5, "text": "<p>This is not an answer, just a reformulation into something which looks perhaps tractable (along the lines of the comment by G. Paseman). Define, for $m\\geq 1$ and $x\\in (0,1)$, $$g_m(x):=\\frac{d^2}{dx^2}[\\log(1+x^m)]=\\frac{x^{m-2}(m(m-1)-mx^m)}{(1+x^m)^2}$$ and $$h_m(x):=\\frac{d^2}{dx^2}[\\log(1-x^m)]=-\\frac{x^{m-2}(m(m-1)+mx^m)}{(1-x^m)^2}.$$ Then, for $n\\geq 2$, $$F_n''(x)=g_{4n-1}(x)+g_{2n}(x)+h_{2n+1}(x)-g_{2n+1}(x)-h_{2n+2}(x).$$ </p>\n\n<p>Two observations</p>\n\n<ul>\n<li>The contribution of the first, second, and last terms is non-negative for all $n\\geq 2$ and $x\\in(0,1)$, while that of the remaining terms is non-positive.</li>\n<li>It would therefore suffice to show the following: for $n$ and $x$ as above,\n$$g_{2n+1}(x)-h_{2n+1}(x)\\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{2n+2}(x).$$</li>\n</ul>\n\n<p><strong>Added</strong>: Since I apparently cannot add a comment, I would just note that while it is certainly (always) possible I overlook something, I have just double checked the algebra with maple, and believe that the expressions are correct (there doesn't seem to be a missing factor of m). Another note (in addition to the symmetry observed in the comment by Paseman) is the (trivial) identity $h_{2m}=g_m+h_m$, leading to another form for the desired bound:\n$$g_{2n+1}(x)+g_{n+1}(x)-h_{2n+1}(x)\\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{n+1}(x).$$</p>\n" }, { "answer_id": 259508, "score": 3, "text": "<p>This is only a comment to @DimaPasechnik, but I cannot put the picture in a comment. The surface to the right of the $x=y$ is a plot of Dima's function (barring mistakes); clearly not convex.</p>\n\n<p><a href=\"https://i.stack.imgur.com/7DSRU.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/7DSRU.jpg\" alt=\"enter image description here\"></a></p>\n" }, { "answer_id": 259489, "score": 2, "text": "<p>I conjecture a stronger statement: that the function $F(x,y)$ of two variables $x,y$ obtained from $F_n(x)$ by substituting $y$ for $x^{2n}$ is convex for $0&lt;y&lt;x&lt;0$. </p>\n\n<p>EDIT2: this conjecture is wrong as stated. See e.g. the plot in the answer by Yaakov Baruch, or rotate the plot in Sage...</p>\n\n<p>While this reduces to checking positive definiteness of \na matrix of bivariate polynomials of degree about 20, obtained from the Hessian of $F$ by clearing common (positive) denominator, this should be possible to make to work by standard real algebraic geometry tools.\nHere is an (ugly) plot of $F$:</p>\n\n<p><a href=\"https://i.stack.imgur.com/l9j8a.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/l9j8a.jpg\" alt=\"plot on [0,1]x[0,1]\"></a> obtained by exporting the plot from jmol in Sage(math) </p>\n\n<pre><code>sage: var('x y')\nsage: F(x,y)=log((x+y^2)*(1+y)*(1-x*y)/(x*(1+x*y)*(1-x^2*y)))\nsage: plot3d(F,[0,1],[0,1])\n</code></pre>\n\n<p>EDIT: comments say that as $y\\to 0+$, the Hessian becomes indefinite. That is, for my original claim to hold, one need to assume $y&gt;\\epsilon&gt;0$, with $\\epsilon$ possibly depending upon $x$. </p>\n" }, { "answer_id": 259370, "score": 21, "text": "<p>Fix $n$, and let $V(x)$ be the polynomial from the question, for which we want to show that it is\npositive on the open interval $(0,1)$. Set\n\\begin{align*}\na(n) &amp;= 1024n^2 - 4096/3n + 320\\\\\nb(n) &amp;= 12288n^3 - 14336n^2 + 4864/3n + 256\\\\\nW(x) &amp;= x^4\\left(V(x)-(n+1)^2(2n+1)^2x^{24n-1}(1-x)^4(a(n)x + b(n)(1-x))\\right).\n\\end{align*}</p>\n\n<p>For all $n\\ge12$ we claim that $(1-x)^6$ divides $W(x)$, and that all the coefficients of $W(x)/(1-x)^6$ are nonnegative. In particular, $W(x)$ is positive for $x&gt;0$. As $a(n)$ and $b(n)$ is positive for all $n\\ge1$, we see that $V(x)$ is positive for all $0&lt;x&lt;1$, provided that $n\\ge12$. The cases $2\\le n\\le 11$ are easily checked directly, for instance by noting that the coefficients of $(1+x)^{\\deg V}V(1/(1+x))$ are positive.</p>\n\n<p>I'm not sure about the easiest kind to prove the assertion about the coefficients of $W(x)/(1-x)^6$. We compute them via\n\\begin{equation}\n\\frac{W(x)}{(1-x)^6}=W(x)\\sum_{k\\ge0}\\binom{k+5}{5}x^k.\n\\end{equation}</p>\n\n<p>All what follows is assisted/confirmed by the Sage code below.</p>\n\n<p>The exponents of $W(x)$ have the form $2ni+j$ for $0\\le i\\le 12$ and $0\\le j\\le 10$. Let $a_{i,j}$ be the corresponding coefficient. It is a polynomial in $n$.</p>\n\n<p>We have\n\\begin{equation} \n\\frac{W(x)}{(1-x)^6} = \\sum_{i,j}a_{i,j}(n)x^{2ni+j}\\sum_k\\binom{k+5}{5}x^k.\n\\end{equation}</p>\n\n<p>With $r\\ge0$ and $0\\le s\\le 2n-1$, the coefficient of $x^{2nr+s}$ in $W(x)\\frac{1}{(1-x)^6}$\nis the sum of all $a_{i,j}(n)\\cdot\\binom{nr+s-ni-j+y}{5}$ for all pairs $(i,j)$ where $i&lt;r$ and $0\\le j\\le 10$, or $i=r$ and $j\\le s$.</p>\n\n<p>We distinguish the cases $0\\le s\\le 9$ from the cases $s\\ge10$. In these latter cases, we write $s=10+y$. Recall that $s\\le 2n-1$, so $10+y=2n-1-o$ for a nonnegative $o$. Upon replacing $n$ with $(11+y+1)/2$, we get polynomials in $n$ and $o$.</p>\n\n<p>In the former cases, the coefficients are polynomials in $n$, which have positive coefficients upon replacing $n$ with $n+12$. So these coefficients are positive for all $n\\ge12$. </p>\n\n<p>In the latter cases it turns out that in all but one case the coefficients are positive. So for nonnegative $y$ and $o$, the values are nonnegative.</p>\n\n<p>For the single exception we see that if we multiply it with a suitable polynomial, the resulting coefficients are positive.</p>\n\n<p><em>Remark (answering Jason's question):</em> It is a known fact that a polynomial $V$ is positive on $(0,1)$ if and only if it is a nonnegative linear combination of polynomials $x^i(1-x)^j$. The problem is that it may involve terms where $i+j$ is bigger than $\\deg V$. (This doesn't happen here, though.) I used an LP solver to play a little with $V$ and $\\frac{V}{(1-x^2)^4x^{2n-2}}$. From that a pattern showed up which lead to a solution.</p>\n\n<p>By the way, $V$ actually seems to be a nonnegative linear combination of $x^i(1-x)^{\\deg V-i}$. This is equivalent to say that all the coefficients of $(1+x)^{\\deg V}V(1/(1+x))$ are nonnegative. It is not hard to compute explicit expressions for these coefficients, but I don't see an easy arguments why they can't be negative.</p>\n\n<pre><code># Formally compute W\nvar('X N')\nP = (1+X^(4*N-1))*(1+X^(2*N))*(1-X^(2*N+1))\nQ = (1+X^(2*N+1))*(1-X^(2*N+2))\nV = P*Q^2*P.diff(X,2)+(P*Q.diff(X))^2-P^2*Q*Q.diff(X,2)-(P.diff(X)*Q)^2\na = 1024*N^2 - 4096/3*N + 320\nb = 12288*N^3 - 14336*N^2 + 4864/3*N + 256\nW = X^4*(V - (2*N + 1)^2 *(N + 1)^2*X^(24*N-1)*(1-X)^4*(a*X + b*(1-X)))\n\n# Check that (1-X)^6 divides W(X)\nprint all(W.diff(X,i)(X=1).polynomial(QQ) == 0 for i in [0..5])\n\n# Compute the coefficients a_ij for the exponents 2ni+j of W. Somewhat\n# clumsy, as I don't know how to deal with polynomials where exponents\n# are symbolic expressions.\n#\n# l is the list of summands of W\nl = [z.canonicalize_radical() for z in W.expand().operands()]\nK.&lt;n&gt; = QQ[]\naij = {(i,j):K(0) for i in [0..12] for j in [0..10]}\nfor term in l:\n c = term(X=1) # get coefficient of term\n e = (term.diff(X)/c)(X=1) # get exponent of term\n c = K(c.polynomial(QQ)) # convert c to proper polynomial in n\n i, j = ZZ(e.diff(N))//2, ZZ(e(N=0)) # Clumsy method to compute pairs (i,j)\n aij[i,j] += c\n\n# Check if coefficients aij[i,j] were correctly computed\nWnew = sum(c(n=N)*X^(2*N*i+j) for (i,j),c in aij.items())\nprint (W-Wnew).canonicalize_radical().is_trivial_zero()\n\ndef bino(k): # binomial coefficient binom(-6,k)=binom(k+5,5)\n return prod(k+5-z for z in range(5))/120\n\n# compute the coefficients of W(X)/(1-X)^6\nK = K.extend_variables(('y', 'o'))\nK.inject_variables()\nfor r in [0..12]:\n for s in [0..10]:\n d = 1 if s == 10 else 0\n # compute coefficient of X^(2nr+s+d*y)\n f = sum(aij[i,j]*bino(2*n*(r-i)+s+d*y-j) for i in [0..r]\n for j in [0..s if i == r else 10])\n # check non-negativity of the polynomial f\n if d == 0:\n # Checks if the coefficient of X^(2nr+s), which is a\n # polynomial in n, has nonnegative coefficients upon replacing\n # n with n+12\n f = f(n=n+12)\n if min(f.coefficients()+[0]) &lt; 0:\n print \"False\"\n else:\n # The coefficient of X^(2nr+10+y), which is a polynomial in n\n # and y, has to be nonnegative for 0 &lt;= 10+y &lt;= 2n-1, so 10+y\n # = 2n-1-o for non-negative o. So upon replacing n with\n # (11+y+o)/2, we get a polynomial in y and o which has to be\n # nonnegative for all nonnegative y and o. In all but one\n # case, this holds because the coeffcients are non-negative.\n f = f(n=(11+y+o)/2)\n if min(f.coefficients()+[0]) &lt; 0:\n c = o^2 + 23*o*y + 1360*y^2 + 99*o + 340*y + 1675\n print min(c.coefficients()+(c*f).coefficients()) &gt;= 0\n</code></pre>\n" }, { "answer_id": 259280, "score": 11, "text": "<p>This proof is mainly based on some of the suggestions in the early version of this answer. Also, we are building here on the comment by Peter Mueller concerning the polynomial $a$ defined below. It turns out that similar techniques can be used for all other nontrivial polynomials arising in the proof, which allowed us to replace multiple uses of Mathematica command Reduce[] in the last version of this answer by showing that the coefficients of the mentioned relevant polynomials are all nonnegative. I have still retained some use of Reduce[] -- but only to quickly make the routine check for $n\\le11$. </p>\n\n<p>Note that \n\\begin{equation}\nG:=\nF_n''(x)\tx^{2-2 n} \\left(x^{2 n}+1\\right)^2 \\left(x^{4 n}+x\\right)^2 \\left(x^{2 n+1}-1\\right)^2\n \\left(x^{2 n+1}+1\\right)^2 \\left(x^{2 n+2}-1\\right)^2\n\\end{equation}\nis a polynomial in $n$, $x$, and $y:=x^{2n}$, of degree $2$ in $n$:\n\\begin{equation}\n\tG=a+bn+cn^2,\n\\end{equation}\nwhere $a,b,c$ are certain polynomials in $x,y$. </p>\n\n<p>The check of $G\\ge0$ for $n=2,\\dots,11$ is straightforward. So, assume $n\\ge12$, whence $0&lt;y&lt;x^{24}&lt;x^{12}$. \nWe also always assume that $0&lt;x&lt;1$, $0&lt;y&lt;x^{24}$ , \n\\begin{equation}\n\tn=\\frac{\\ln y}{2\\ln x}; \\tag{1}\n\\end{equation}\nthe latter relation is of course just another form of the equality $y=x^{2n}$. </p>\n\n<p>Note that \n$0&lt;y&lt;1\\ \\&amp;\\ 0&lt;x&lt;1$ implies $x=1/(1+u)\\ \\&amp;\\ y=1/(1+v)$ for some $u,v\\ge0$. \nAs suggested by Peter Mueller, it turns out that \n\\begin{equation}\n\ta\\; (1 + u)^{10} (1 + v)^{11}|_{x=1/(1 + u), y=1/(1 + v)}\n\\end{equation}\nis a polynomial in $u,v$ with all coefficients nonnegative. \nDetails of all calculations can be seen in the <a href=\"https://www.dropbox.com/s/g87a59uzdsb75m5/SolutionM1.nb?dl=0\" rel=\"nofollow noreferrer\">Mathematica notebook</a> \nand its <a href=\"https://www.dropbox.com/s/k4vlx4brvok099k/SolutionM1.pdf?dl=0\" rel=\"nofollow noreferrer\">pdf image</a></p>\n\n<p>So, \n\\begin{equation}\n\ta\\ge0\n\\end{equation}\nfor all $x,y$ in $(0,1)$. </p>\n\n<p>Next, introduce \n\\begin{equation}\nx_* := \\frac{869}{1000},\\quad \nx_0:=\\frac{7655}{10000},\\quad y_0:= x_0^{24}\\approx0.00164,\\quad\nx_1 := \\frac{985}{1000},\\quad y_1:= x_1^{24}\\approx0.696. \n\\end{equation} </p>\n\n<p>The following cases/subcases/subsubcases are exhaustive: </p>\n\n<p><strong>Case 1: $y\\le y_1$.</strong> Then the condition $0&lt;y\\le y_1\\ \\&amp;\\ 0&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;x^{24}$ implies $x=x_1/(1+u)\\ \\&amp;\\ y=x^{24}/(1+v)$ or $x=x_1+(1-x_1)/(1+u)\\ \\&amp;\\ y=y_1/(1+v)$ for some $u\\ge0$ and $v\\ge0$, depending on whether $x&lt;x_1$ or $x\\ge x_1$. \nIt turns out that \n\\begin{equation}\n\tc\\; (1 + u)^{248} (1 + v)^{10}|_{x=x_1/(1+u)\\ \\&amp;\\ y=x^{24}/(1+v)}\n\\end{equation}\nand \n\\begin{equation}\n\tc\\; (1 + u)^{10} (1 + v)^{10}|_{x=x_1+(1-x_1)/(1+u)\\ \\&amp;\\ y=y_1/(1+v)}\n\\end{equation}\nare each a polynomial in $u,v$ with all coefficients nonnegative. \nSo, \n\\begin{equation}\n\tc\\ge0 \\quad\\text{in Case 1}. \n\\end{equation}\nSo, for the derivative $G'_n$ of $G$ in $n$, we have \n\\begin{equation}\\tag{2}\nG'_n=b+2cn\\ge b+24c=(G'_n)|_{n=12}.\t\n\\end{equation}</p>\n\n<p>Now we have to distinguish two subcases of Case 1, with further subsubcases of Subcase 1.1: </p>\n\n<p><strong>Subcase 1.1: $y\\le y_1$ and ($y\\ge y_0$ or $x\\le x_*$).</strong> </p>\n\n<p><strong>Subsubcase 1.1.1: $y\\le y_1$ and $y\\ge y_0$ and $x\\ge x_1$.</strong> Then $y\\le y_1\\ \\&amp;\\ y\\ge y_0\\ \\&amp;\\ x\\ge x_1 \\ \\&amp;\\ 0&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;x^{24}$ implies \n$$x=x_{111}:=x_1+(1-x_1)/(1+u)\\ \\&amp;\\ y=y_{111}:=y_0+(y_1-y_0)/(1+v)$$\nfor some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=12}\\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=12}\\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p>\n\n<p><strong>Subsubcase 1.1.2: $y\\le y_1$ and $y\\ge y_0$ and $x&lt;x_1$.</strong> Then $x=x_{112}:=x_0+(x_1-x_0)/(1+u)\\ \\&amp;\\ y=y_{112}:=y_0+(x^{24}-y_0)/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112}, y=y_{112}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112},y=y_{112}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p>\n\n<p><strong>Subsubcase 1.1.3: $y\\le y_1$ and $x\\le x_*$.</strong> Then $x=x_{113}:=x_*/(1+u)\\ \\&amp;\\ y=y_{113}:=x^{24}/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p>\n\n<p>So, $G|_{n=12}\\ge0$ and $(G'_n)|_{n=12}\\ge0$ in all the three subsubcases of Subcase 1.1. In view of $(2)$, we have \n$G\\ge0$ in Subcase 1.1. </p>\n\n<p><strong>Subcase 1.2: $y\\le y_1$ and $y&lt;y_0$ and $x&gt;x_*$.</strong> Then, of course, the condition $y\\le y_1$ is redundant. Let here \n\\begin{equation}\n\t\\rho(x):=\\frac7{2(1-x)}. \n\\end{equation}\nThe condition \n$x_*&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;y_0$ implies $x=x_{12}:=x_*+(1-x_*)/(1+u)$ and $y=y_{12}:=y_0/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=\\rho(x)/2}\\; u\\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=\\rho(x)/2}\\; u^2\\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. \nSo, similarly to Subcase 1.1, for all $n\\ge\\rho(x)/2$ we get $G\\ge G|_{n=\\rho(x)/2}\\ge0$. \nIn Subcase 1.2, it remains to note that for our particular $n$, as in $(1)$, we indeed have $n\\ge\\rho(x)/2$. This follows because for $y&lt;y_0$ and $x&gt;x_*$ \n\\begin{equation}\n\t\\ln\\frac1y&gt;\\ln\\frac1{y_0}&gt;1.08\\frac7{2(1-x_*)}\\,\\ln\\frac1{x_*}\n\t&gt;\\frac7{2(1-x)}\\,\\ln\\frac1{x}=\\rho(x)\\ln\\frac1{x}. \n\\end{equation}\nSo, $G\\ge0$ in Subcase 1.2 as well. </p>\n\n<p>It remains to consider </p>\n\n<p><strong>Case 2: $y&gt;y_1$.</strong> Then condition $y&gt;y_1\\ \\&amp;\\ 0&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;x^{24}[&lt;x^{12}]$ implies $x=x_2:=x_1+(1-x_1)/(1+u),y=y_2:=y_1+(x^{12}-y_1)/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\tb\\;(1 + u)^{140} (1 + v)^{11}|_{x=x_{2},y=y_{2}}\n\\end{equation}\nis also a polynomial in $u,v$ with all coefficients nonnegative. \nHence, $b\\ge0$ in Case 2. \nSo, if $c\\ge0$, then trivially $G\\ge0$ (since $a\\ge0$ always). So, without loss of generality, $c&lt;0$ and hence $G$ is concave in $n$. So, it is enough to bracket the $n$ as in $(1)$ between some simple rational expressions $n_1$ and $n_2$ such that $G|_{n=n_1}\\ge0$ and $G|_{n=n_2}\\ge0$. </p>\n\n<p>In Case 2, $y$ is \"close\" to $1$ and hence, in view of inequality $y&lt;x^{24}$, so is $x$: $x&gt;y_1^{1/12}=x_1=0.985$. So, it is a bit more convenient here to change variables $x, y$ to \"small variables\" $u:=1-x$ and $v:=1-y$. \nThen $0&lt;v&lt;1-y_1\\approx0.304$ and $1-u=x&gt;y^{1/24}=(1-v)^{1/24}&gt;1-v/3$, whence $u&lt;v/3$. Let now $t:=u/v$, so that $0&lt;t&lt;1/3$. \nIt follows, in Case 2, that $t=(1/3)/(1+r)$ and $v=(1-y_1)/(1+s)$ for some $r\\ge0$ and $s\\ge0$. </p>\n\n<p>The mentioned brackets $n_1$ and $n_2$ are respectively $\\ell_1/2$ and $\\ell_2/2$, where<br>\n\\begin{equation}\n\t\\ell_1:=\\frac1t,\\quad\\ell_2:=\\frac{(2 - v) (1 - t v)}{t (1 - v) (2 - t v)}. \n\\end{equation}\nIt is not hard to see (details on this are in the same Mathematica notebook) that then indeed $n_1 &lt; n &lt; n_2$. \nIt remains to show that $G|_{n=n_1}\\ge0$ and $G|_{n=n_2}\\ge0$. </p>\n\n<p>It turns out that \n\\begin{equation}\n\t(G'_n)|_{n=\\ell_1/2;\\,x=1-tv,\\,y=1-v;\\,t=(1/3)/(1+r),\\,v=(1-y_1)/(1+s)}\\; (1 + r)^{10} (1 + s)^{19}\n\\end{equation}\nand \n\\begin{equation}\n\t(G'_n)|_{n=\\ell_2/2;\\,x=1-tv,\\,y=1-v;\\,t=(1/3)/(1+r),\\,v=(1-y_1)/(1+s)}\\; (1 + r)^{10} (1 + s)^{19}\\; q\n\\end{equation}\nare each a polynomial in $r,s$ with all coefficients nonnegative, \nwhere \n\\begin{multline*}\n\tq:=\n\t(1 + r)^{10} (1 + s)^{19} \\\\\n\t\\times(11673186598630578538556565100133681446610566511878526881 \\\\ \n\t+ \n 16777216000000000000000000000000000000000000000000000000 s)^2 \\\\ \n\\times\\big(31853088866210192846185521700044560482203522170626175627 \\\\\n+ \n 33554432000000000000000000000000000000000000000000000000 (r + s + \n r s)\\big)^2. \n\\end{multline*}</p>\n\n<p>The proof is complete. </p>\n" }, { "answer_id": 259260, "score": 10, "text": "<p>I sketch a method that should show $F_n$ is convex for $n$ sufficiently large.</p>\n\n<p>A Taylor expansion gives that $F_n(x) = x^{2n} (1-x)^2 + O(x^{4n-1})$, and so $$F_n''(x) = 2x^{2n-2}[2n^2(1-x)^2 + x^2 - n(1-x)(1+3x)] + O(x^{4n-3}).$$\nThis should show that $F_n''(x) &gt; 0$ for $n$ large, provided $x &lt; 1-\\frac{C}{n}$ for some fixed $C&gt;0$.</p>\n\n<p>On the other hand, according to Mathematica, $F_n''(1) = 4n^2 - \\frac{16}{3} n + \\frac{5}{4}$, which has its largest root at $\\frac{8 + \\sqrt{19}}{12} \\approx 1.02991$. (By the way, there doesn't seem to be any reason that $n$ should be restricted to be an integer, and this is in line with some plots I made indicating $F_n(x)$ is convex for $n &gt; 1.03$.)\nMoreover, Mathematica gives $$\\lim_{n\\rightarrow \\infty} n^{-2}F_n''(1-\\frac{y}{n}) = \\frac{16 e^{4y}}{(1+e^{4y})^2} &gt; 0.$$</p>\n\n<p>Putting these two pieces together should be able to show $F_n(x)$ is convex on $(0,1)$ for $n$ sufficiently large. Unfortunately, it might be very painful to work out all the error terms explicitly.</p>\n" }, { "answer_id": 259207, "score": 6, "text": "<p>Not an answer, just a plot of $F_n(x)$ for $n=2,\\ldots,10$:\n<hr />\n&nbsp; &nbsp; &nbsp; &nbsp; &nbsp;\n<a href=\"https://i.stack.imgur.com/jzWwY.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/jzWwY.jpg\" alt=\"Fn(x)\"></a>\n<hr />\n&nbsp; &nbsp; &nbsp; &nbsp; &nbsp;\n<a href=\"https://i.stack.imgur.com/c8gFL.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/c8gFL.jpg\" alt=\"Fn(x)2\"></a></p>\n\n<hr />\n" }, { "answer_id": 259213, "score": 5, "text": "<p>This is not an answer, just a reformulation into something which looks perhaps tractable (along the lines of the comment by G. Paseman). Define, for $m\\geq 1$ and $x\\in (0,1)$, $$g_m(x):=\\frac{d^2}{dx^2}[\\log(1+x^m)]=\\frac{x^{m-2}(m(m-1)-mx^m)}{(1+x^m)^2}$$ and $$h_m(x):=\\frac{d^2}{dx^2}[\\log(1-x^m)]=-\\frac{x^{m-2}(m(m-1)+mx^m)}{(1-x^m)^2}.$$ Then, for $n\\geq 2$, $$F_n''(x)=g_{4n-1}(x)+g_{2n}(x)+h_{2n+1}(x)-g_{2n+1}(x)-h_{2n+2}(x).$$ </p>\n\n<p>Two observations</p>\n\n<ul>\n<li>The contribution of the first, second, and last terms is non-negative for all $n\\geq 2$ and $x\\in(0,1)$, while that of the remaining terms is non-positive.</li>\n<li>It would therefore suffice to show the following: for $n$ and $x$ as above,\n$$g_{2n+1}(x)-h_{2n+1}(x)\\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{2n+2}(x).$$</li>\n</ul>\n\n<p><strong>Added</strong>: Since I apparently cannot add a comment, I would just note that while it is certainly (always) possible I overlook something, I have just double checked the algebra with maple, and believe that the expressions are correct (there doesn't seem to be a missing factor of m). Another note (in addition to the symmetry observed in the comment by Paseman) is the (trivial) identity $h_{2m}=g_m+h_m$, leading to another form for the desired bound:\n$$g_{2n+1}(x)+g_{n+1}(x)-h_{2n+1}(x)\\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{n+1}(x).$$</p>\n" }, { "answer_id": 259508, "score": 3, "text": "<p>This is only a comment to @DimaPasechnik, but I cannot put the picture in a comment. The surface to the right of the $x=y$ is a plot of Dima's function (barring mistakes); clearly not convex.</p>\n\n<p><a href=\"https://i.stack.imgur.com/7DSRU.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/7DSRU.jpg\" alt=\"enter image description here\"></a></p>\n" }, { "answer_id": 259489, "score": 2, "text": "<p>I conjecture a stronger statement: that the function $F(x,y)$ of two variables $x,y$ obtained from $F_n(x)$ by substituting $y$ for $x^{2n}$ is convex for $0&lt;y&lt;x&lt;0$. </p>\n\n<p>EDIT2: this conjecture is wrong as stated. See e.g. the plot in the answer by Yaakov Baruch, or rotate the plot in Sage...</p>\n\n<p>While this reduces to checking positive definiteness of \na matrix of bivariate polynomials of degree about 20, obtained from the Hessian of $F$ by clearing common (positive) denominator, this should be possible to make to work by standard real algebraic geometry tools.\nHere is an (ugly) plot of $F$:</p>\n\n<p><a href=\"https://i.stack.imgur.com/l9j8a.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/l9j8a.jpg\" alt=\"plot on [0,1]x[0,1]\"></a> obtained by exporting the plot from jmol in Sage(math) </p>\n\n<pre><code>sage: var('x y')\nsage: F(x,y)=log((x+y^2)*(1+y)*(1-x*y)/(x*(1+x*y)*(1-x^2*y)))\nsage: plot3d(F,[0,1],[0,1])\n</code></pre>\n\n<p>EDIT: comments say that as $y\\to 0+$, the Hessian becomes indefinite. That is, for my original claim to hold, one need to assume $y&gt;\\epsilon&gt;0$, with $\\epsilon$ possibly depending upon $x$. </p>\n" }, { "answer_id": 259370, "score": 21, "text": "<p>Fix $n$, and let $V(x)$ be the polynomial from the question, for which we want to show that it is\npositive on the open interval $(0,1)$. Set\n\\begin{align*}\na(n) &amp;= 1024n^2 - 4096/3n + 320\\\\\nb(n) &amp;= 12288n^3 - 14336n^2 + 4864/3n + 256\\\\\nW(x) &amp;= x^4\\left(V(x)-(n+1)^2(2n+1)^2x^{24n-1}(1-x)^4(a(n)x + b(n)(1-x))\\right).\n\\end{align*}</p>\n\n<p>For all $n\\ge12$ we claim that $(1-x)^6$ divides $W(x)$, and that all the coefficients of $W(x)/(1-x)^6$ are nonnegative. In particular, $W(x)$ is positive for $x&gt;0$. As $a(n)$ and $b(n)$ is positive for all $n\\ge1$, we see that $V(x)$ is positive for all $0&lt;x&lt;1$, provided that $n\\ge12$. The cases $2\\le n\\le 11$ are easily checked directly, for instance by noting that the coefficients of $(1+x)^{\\deg V}V(1/(1+x))$ are positive.</p>\n\n<p>I'm not sure about the easiest kind to prove the assertion about the coefficients of $W(x)/(1-x)^6$. We compute them via\n\\begin{equation}\n\\frac{W(x)}{(1-x)^6}=W(x)\\sum_{k\\ge0}\\binom{k+5}{5}x^k.\n\\end{equation}</p>\n\n<p>All what follows is assisted/confirmed by the Sage code below.</p>\n\n<p>The exponents of $W(x)$ have the form $2ni+j$ for $0\\le i\\le 12$ and $0\\le j\\le 10$. Let $a_{i,j}$ be the corresponding coefficient. It is a polynomial in $n$.</p>\n\n<p>We have\n\\begin{equation} \n\\frac{W(x)}{(1-x)^6} = \\sum_{i,j}a_{i,j}(n)x^{2ni+j}\\sum_k\\binom{k+5}{5}x^k.\n\\end{equation}</p>\n\n<p>With $r\\ge0$ and $0\\le s\\le 2n-1$, the coefficient of $x^{2nr+s}$ in $W(x)\\frac{1}{(1-x)^6}$\nis the sum of all $a_{i,j}(n)\\cdot\\binom{nr+s-ni-j+y}{5}$ for all pairs $(i,j)$ where $i&lt;r$ and $0\\le j\\le 10$, or $i=r$ and $j\\le s$.</p>\n\n<p>We distinguish the cases $0\\le s\\le 9$ from the cases $s\\ge10$. In these latter cases, we write $s=10+y$. Recall that $s\\le 2n-1$, so $10+y=2n-1-o$ for a nonnegative $o$. Upon replacing $n$ with $(11+y+1)/2$, we get polynomials in $n$ and $o$.</p>\n\n<p>In the former cases, the coefficients are polynomials in $n$, which have positive coefficients upon replacing $n$ with $n+12$. So these coefficients are positive for all $n\\ge12$. </p>\n\n<p>In the latter cases it turns out that in all but one case the coefficients are positive. So for nonnegative $y$ and $o$, the values are nonnegative.</p>\n\n<p>For the single exception we see that if we multiply it with a suitable polynomial, the resulting coefficients are positive.</p>\n\n<p><em>Remark (answering Jason's question):</em> It is a known fact that a polynomial $V$ is positive on $(0,1)$ if and only if it is a nonnegative linear combination of polynomials $x^i(1-x)^j$. The problem is that it may involve terms where $i+j$ is bigger than $\\deg V$. (This doesn't happen here, though.) I used an LP solver to play a little with $V$ and $\\frac{V}{(1-x^2)^4x^{2n-2}}$. From that a pattern showed up which lead to a solution.</p>\n\n<p>By the way, $V$ actually seems to be a nonnegative linear combination of $x^i(1-x)^{\\deg V-i}$. This is equivalent to say that all the coefficients of $(1+x)^{\\deg V}V(1/(1+x))$ are nonnegative. It is not hard to compute explicit expressions for these coefficients, but I don't see an easy arguments why they can't be negative.</p>\n\n<pre><code># Formally compute W\nvar('X N')\nP = (1+X^(4*N-1))*(1+X^(2*N))*(1-X^(2*N+1))\nQ = (1+X^(2*N+1))*(1-X^(2*N+2))\nV = P*Q^2*P.diff(X,2)+(P*Q.diff(X))^2-P^2*Q*Q.diff(X,2)-(P.diff(X)*Q)^2\na = 1024*N^2 - 4096/3*N + 320\nb = 12288*N^3 - 14336*N^2 + 4864/3*N + 256\nW = X^4*(V - (2*N + 1)^2 *(N + 1)^2*X^(24*N-1)*(1-X)^4*(a*X + b*(1-X)))\n\n# Check that (1-X)^6 divides W(X)\nprint all(W.diff(X,i)(X=1).polynomial(QQ) == 0 for i in [0..5])\n\n# Compute the coefficients a_ij for the exponents 2ni+j of W. Somewhat\n# clumsy, as I don't know how to deal with polynomials where exponents\n# are symbolic expressions.\n#\n# l is the list of summands of W\nl = [z.canonicalize_radical() for z in W.expand().operands()]\nK.&lt;n&gt; = QQ[]\naij = {(i,j):K(0) for i in [0..12] for j in [0..10]}\nfor term in l:\n c = term(X=1) # get coefficient of term\n e = (term.diff(X)/c)(X=1) # get exponent of term\n c = K(c.polynomial(QQ)) # convert c to proper polynomial in n\n i, j = ZZ(e.diff(N))//2, ZZ(e(N=0)) # Clumsy method to compute pairs (i,j)\n aij[i,j] += c\n\n# Check if coefficients aij[i,j] were correctly computed\nWnew = sum(c(n=N)*X^(2*N*i+j) for (i,j),c in aij.items())\nprint (W-Wnew).canonicalize_radical().is_trivial_zero()\n\ndef bino(k): # binomial coefficient binom(-6,k)=binom(k+5,5)\n return prod(k+5-z for z in range(5))/120\n\n# compute the coefficients of W(X)/(1-X)^6\nK = K.extend_variables(('y', 'o'))\nK.inject_variables()\nfor r in [0..12]:\n for s in [0..10]:\n d = 1 if s == 10 else 0\n # compute coefficient of X^(2nr+s+d*y)\n f = sum(aij[i,j]*bino(2*n*(r-i)+s+d*y-j) for i in [0..r]\n for j in [0..s if i == r else 10])\n # check non-negativity of the polynomial f\n if d == 0:\n # Checks if the coefficient of X^(2nr+s), which is a\n # polynomial in n, has nonnegative coefficients upon replacing\n # n with n+12\n f = f(n=n+12)\n if min(f.coefficients()+[0]) &lt; 0:\n print \"False\"\n else:\n # The coefficient of X^(2nr+10+y), which is a polynomial in n\n # and y, has to be nonnegative for 0 &lt;= 10+y &lt;= 2n-1, so 10+y\n # = 2n-1-o for non-negative o. So upon replacing n with\n # (11+y+o)/2, we get a polynomial in y and o which has to be\n # nonnegative for all nonnegative y and o. In all but one\n # case, this holds because the coeffcients are non-negative.\n f = f(n=(11+y+o)/2)\n if min(f.coefficients()+[0]) &lt; 0:\n c = o^2 + 23*o*y + 1360*y^2 + 99*o + 340*y + 1675\n print min(c.coefficients()+(c*f).coefficients()) &gt;= 0\n</code></pre>\n" }, { "answer_id": 259280, "score": 11, "text": "<p>This proof is mainly based on some of the suggestions in the early version of this answer. Also, we are building here on the comment by Peter Mueller concerning the polynomial $a$ defined below. It turns out that similar techniques can be used for all other nontrivial polynomials arising in the proof, which allowed us to replace multiple uses of Mathematica command Reduce[] in the last version of this answer by showing that the coefficients of the mentioned relevant polynomials are all nonnegative. I have still retained some use of Reduce[] -- but only to quickly make the routine check for $n\\le11$. </p>\n\n<p>Note that \n\\begin{equation}\nG:=\nF_n''(x)\tx^{2-2 n} \\left(x^{2 n}+1\\right)^2 \\left(x^{4 n}+x\\right)^2 \\left(x^{2 n+1}-1\\right)^2\n \\left(x^{2 n+1}+1\\right)^2 \\left(x^{2 n+2}-1\\right)^2\n\\end{equation}\nis a polynomial in $n$, $x$, and $y:=x^{2n}$, of degree $2$ in $n$:\n\\begin{equation}\n\tG=a+bn+cn^2,\n\\end{equation}\nwhere $a,b,c$ are certain polynomials in $x,y$. </p>\n\n<p>The check of $G\\ge0$ for $n=2,\\dots,11$ is straightforward. So, assume $n\\ge12$, whence $0&lt;y&lt;x^{24}&lt;x^{12}$. \nWe also always assume that $0&lt;x&lt;1$, $0&lt;y&lt;x^{24}$ , \n\\begin{equation}\n\tn=\\frac{\\ln y}{2\\ln x}; \\tag{1}\n\\end{equation}\nthe latter relation is of course just another form of the equality $y=x^{2n}$. </p>\n\n<p>Note that \n$0&lt;y&lt;1\\ \\&amp;\\ 0&lt;x&lt;1$ implies $x=1/(1+u)\\ \\&amp;\\ y=1/(1+v)$ for some $u,v\\ge0$. \nAs suggested by Peter Mueller, it turns out that \n\\begin{equation}\n\ta\\; (1 + u)^{10} (1 + v)^{11}|_{x=1/(1 + u), y=1/(1 + v)}\n\\end{equation}\nis a polynomial in $u,v$ with all coefficients nonnegative. \nDetails of all calculations can be seen in the <a href=\"https://www.dropbox.com/s/g87a59uzdsb75m5/SolutionM1.nb?dl=0\" rel=\"nofollow noreferrer\">Mathematica notebook</a> \nand its <a href=\"https://www.dropbox.com/s/k4vlx4brvok099k/SolutionM1.pdf?dl=0\" rel=\"nofollow noreferrer\">pdf image</a></p>\n\n<p>So, \n\\begin{equation}\n\ta\\ge0\n\\end{equation}\nfor all $x,y$ in $(0,1)$. </p>\n\n<p>Next, introduce \n\\begin{equation}\nx_* := \\frac{869}{1000},\\quad \nx_0:=\\frac{7655}{10000},\\quad y_0:= x_0^{24}\\approx0.00164,\\quad\nx_1 := \\frac{985}{1000},\\quad y_1:= x_1^{24}\\approx0.696. \n\\end{equation} </p>\n\n<p>The following cases/subcases/subsubcases are exhaustive: </p>\n\n<p><strong>Case 1: $y\\le y_1$.</strong> Then the condition $0&lt;y\\le y_1\\ \\&amp;\\ 0&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;x^{24}$ implies $x=x_1/(1+u)\\ \\&amp;\\ y=x^{24}/(1+v)$ or $x=x_1+(1-x_1)/(1+u)\\ \\&amp;\\ y=y_1/(1+v)$ for some $u\\ge0$ and $v\\ge0$, depending on whether $x&lt;x_1$ or $x\\ge x_1$. \nIt turns out that \n\\begin{equation}\n\tc\\; (1 + u)^{248} (1 + v)^{10}|_{x=x_1/(1+u)\\ \\&amp;\\ y=x^{24}/(1+v)}\n\\end{equation}\nand \n\\begin{equation}\n\tc\\; (1 + u)^{10} (1 + v)^{10}|_{x=x_1+(1-x_1)/(1+u)\\ \\&amp;\\ y=y_1/(1+v)}\n\\end{equation}\nare each a polynomial in $u,v$ with all coefficients nonnegative. \nSo, \n\\begin{equation}\n\tc\\ge0 \\quad\\text{in Case 1}. \n\\end{equation}\nSo, for the derivative $G'_n$ of $G$ in $n$, we have \n\\begin{equation}\\tag{2}\nG'_n=b+2cn\\ge b+24c=(G'_n)|_{n=12}.\t\n\\end{equation}</p>\n\n<p>Now we have to distinguish two subcases of Case 1, with further subsubcases of Subcase 1.1: </p>\n\n<p><strong>Subcase 1.1: $y\\le y_1$ and ($y\\ge y_0$ or $x\\le x_*$).</strong> </p>\n\n<p><strong>Subsubcase 1.1.1: $y\\le y_1$ and $y\\ge y_0$ and $x\\ge x_1$.</strong> Then $y\\le y_1\\ \\&amp;\\ y\\ge y_0\\ \\&amp;\\ x\\ge x_1 \\ \\&amp;\\ 0&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;x^{24}$ implies \n$$x=x_{111}:=x_1+(1-x_1)/(1+u)\\ \\&amp;\\ y=y_{111}:=y_0+(y_1-y_0)/(1+v)$$\nfor some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=12}\\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=12}\\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p>\n\n<p><strong>Subsubcase 1.1.2: $y\\le y_1$ and $y\\ge y_0$ and $x&lt;x_1$.</strong> Then $x=x_{112}:=x_0+(x_1-x_0)/(1+u)\\ \\&amp;\\ y=y_{112}:=y_0+(x^{24}-y_0)/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112}, y=y_{112}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112},y=y_{112}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p>\n\n<p><strong>Subsubcase 1.1.3: $y\\le y_1$ and $x\\le x_*$.</strong> Then $x=x_{113}:=x_*/(1+u)\\ \\&amp;\\ y=y_{113}:=x^{24}/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p>\n\n<p>So, $G|_{n=12}\\ge0$ and $(G'_n)|_{n=12}\\ge0$ in all the three subsubcases of Subcase 1.1. In view of $(2)$, we have \n$G\\ge0$ in Subcase 1.1. </p>\n\n<p><strong>Subcase 1.2: $y\\le y_1$ and $y&lt;y_0$ and $x&gt;x_*$.</strong> Then, of course, the condition $y\\le y_1$ is redundant. Let here \n\\begin{equation}\n\t\\rho(x):=\\frac7{2(1-x)}. \n\\end{equation}\nThe condition \n$x_*&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;y_0$ implies $x=x_{12}:=x_*+(1-x_*)/(1+u)$ and $y=y_{12}:=y_0/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=\\rho(x)/2}\\; u\\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=\\rho(x)/2}\\; u^2\\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. \nSo, similarly to Subcase 1.1, for all $n\\ge\\rho(x)/2$ we get $G\\ge G|_{n=\\rho(x)/2}\\ge0$. \nIn Subcase 1.2, it remains to note that for our particular $n$, as in $(1)$, we indeed have $n\\ge\\rho(x)/2$. This follows because for $y&lt;y_0$ and $x&gt;x_*$ \n\\begin{equation}\n\t\\ln\\frac1y&gt;\\ln\\frac1{y_0}&gt;1.08\\frac7{2(1-x_*)}\\,\\ln\\frac1{x_*}\n\t&gt;\\frac7{2(1-x)}\\,\\ln\\frac1{x}=\\rho(x)\\ln\\frac1{x}. \n\\end{equation}\nSo, $G\\ge0$ in Subcase 1.2 as well. </p>\n\n<p>It remains to consider </p>\n\n<p><strong>Case 2: $y&gt;y_1$.</strong> Then condition $y&gt;y_1\\ \\&amp;\\ 0&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;x^{24}[&lt;x^{12}]$ implies $x=x_2:=x_1+(1-x_1)/(1+u),y=y_2:=y_1+(x^{12}-y_1)/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\tb\\;(1 + u)^{140} (1 + v)^{11}|_{x=x_{2},y=y_{2}}\n\\end{equation}\nis also a polynomial in $u,v$ with all coefficients nonnegative. \nHence, $b\\ge0$ in Case 2. \nSo, if $c\\ge0$, then trivially $G\\ge0$ (since $a\\ge0$ always). So, without loss of generality, $c&lt;0$ and hence $G$ is concave in $n$. So, it is enough to bracket the $n$ as in $(1)$ between some simple rational expressions $n_1$ and $n_2$ such that $G|_{n=n_1}\\ge0$ and $G|_{n=n_2}\\ge0$. </p>\n\n<p>In Case 2, $y$ is \"close\" to $1$ and hence, in view of inequality $y&lt;x^{24}$, so is $x$: $x&gt;y_1^{1/12}=x_1=0.985$. So, it is a bit more convenient here to change variables $x, y$ to \"small variables\" $u:=1-x$ and $v:=1-y$. \nThen $0&lt;v&lt;1-y_1\\approx0.304$ and $1-u=x&gt;y^{1/24}=(1-v)^{1/24}&gt;1-v/3$, whence $u&lt;v/3$. Let now $t:=u/v$, so that $0&lt;t&lt;1/3$. \nIt follows, in Case 2, that $t=(1/3)/(1+r)$ and $v=(1-y_1)/(1+s)$ for some $r\\ge0$ and $s\\ge0$. </p>\n\n<p>The mentioned brackets $n_1$ and $n_2$ are respectively $\\ell_1/2$ and $\\ell_2/2$, where<br>\n\\begin{equation}\n\t\\ell_1:=\\frac1t,\\quad\\ell_2:=\\frac{(2 - v) (1 - t v)}{t (1 - v) (2 - t v)}. \n\\end{equation}\nIt is not hard to see (details on this are in the same Mathematica notebook) that then indeed $n_1 &lt; n &lt; n_2$. \nIt remains to show that $G|_{n=n_1}\\ge0$ and $G|_{n=n_2}\\ge0$. </p>\n\n<p>It turns out that \n\\begin{equation}\n\t(G'_n)|_{n=\\ell_1/2;\\,x=1-tv,\\,y=1-v;\\,t=(1/3)/(1+r),\\,v=(1-y_1)/(1+s)}\\; (1 + r)^{10} (1 + s)^{19}\n\\end{equation}\nand \n\\begin{equation}\n\t(G'_n)|_{n=\\ell_2/2;\\,x=1-tv,\\,y=1-v;\\,t=(1/3)/(1+r),\\,v=(1-y_1)/(1+s)}\\; (1 + r)^{10} (1 + s)^{19}\\; q\n\\end{equation}\nare each a polynomial in $r,s$ with all coefficients nonnegative, \nwhere \n\\begin{multline*}\n\tq:=\n\t(1 + r)^{10} (1 + s)^{19} \\\\\n\t\\times(11673186598630578538556565100133681446610566511878526881 \\\\ \n\t+ \n 16777216000000000000000000000000000000000000000000000000 s)^2 \\\\ \n\\times\\big(31853088866210192846185521700044560482203522170626175627 \\\\\n+ \n 33554432000000000000000000000000000000000000000000000000 (r + s + \n r s)\\big)^2. \n\\end{multline*}</p>\n\n<p>The proof is complete. </p>\n" }, { "answer_id": 259260, "score": 10, "text": "<p>I sketch a method that should show $F_n$ is convex for $n$ sufficiently large.</p>\n\n<p>A Taylor expansion gives that $F_n(x) = x^{2n} (1-x)^2 + O(x^{4n-1})$, and so $$F_n''(x) = 2x^{2n-2}[2n^2(1-x)^2 + x^2 - n(1-x)(1+3x)] + O(x^{4n-3}).$$\nThis should show that $F_n''(x) &gt; 0$ for $n$ large, provided $x &lt; 1-\\frac{C}{n}$ for some fixed $C&gt;0$.</p>\n\n<p>On the other hand, according to Mathematica, $F_n''(1) = 4n^2 - \\frac{16}{3} n + \\frac{5}{4}$, which has its largest root at $\\frac{8 + \\sqrt{19}}{12} \\approx 1.02991$. (By the way, there doesn't seem to be any reason that $n$ should be restricted to be an integer, and this is in line with some plots I made indicating $F_n(x)$ is convex for $n &gt; 1.03$.)\nMoreover, Mathematica gives $$\\lim_{n\\rightarrow \\infty} n^{-2}F_n''(1-\\frac{y}{n}) = \\frac{16 e^{4y}}{(1+e^{4y})^2} &gt; 0.$$</p>\n\n<p>Putting these two pieces together should be able to show $F_n(x)$ is convex on $(0,1)$ for $n$ sufficiently large. Unfortunately, it might be very painful to work out all the error terms explicitly.</p>\n" }, { "answer_id": 259207, "score": 6, "text": "<p>Not an answer, just a plot of $F_n(x)$ for $n=2,\\ldots,10$:\n<hr />\n&nbsp; &nbsp; &nbsp; &nbsp; &nbsp;\n<a href=\"https://i.stack.imgur.com/jzWwY.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/jzWwY.jpg\" alt=\"Fn(x)\"></a>\n<hr />\n&nbsp; &nbsp; &nbsp; &nbsp; &nbsp;\n<a href=\"https://i.stack.imgur.com/c8gFL.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/c8gFL.jpg\" alt=\"Fn(x)2\"></a></p>\n\n<hr />\n" }, { "answer_id": 259213, "score": 5, "text": "<p>This is not an answer, just a reformulation into something which looks perhaps tractable (along the lines of the comment by G. Paseman). Define, for $m\\geq 1$ and $x\\in (0,1)$, $$g_m(x):=\\frac{d^2}{dx^2}[\\log(1+x^m)]=\\frac{x^{m-2}(m(m-1)-mx^m)}{(1+x^m)^2}$$ and $$h_m(x):=\\frac{d^2}{dx^2}[\\log(1-x^m)]=-\\frac{x^{m-2}(m(m-1)+mx^m)}{(1-x^m)^2}.$$ Then, for $n\\geq 2$, $$F_n''(x)=g_{4n-1}(x)+g_{2n}(x)+h_{2n+1}(x)-g_{2n+1}(x)-h_{2n+2}(x).$$ </p>\n\n<p>Two observations</p>\n\n<ul>\n<li>The contribution of the first, second, and last terms is non-negative for all $n\\geq 2$ and $x\\in(0,1)$, while that of the remaining terms is non-positive.</li>\n<li>It would therefore suffice to show the following: for $n$ and $x$ as above,\n$$g_{2n+1}(x)-h_{2n+1}(x)\\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{2n+2}(x).$$</li>\n</ul>\n\n<p><strong>Added</strong>: Since I apparently cannot add a comment, I would just note that while it is certainly (always) possible I overlook something, I have just double checked the algebra with maple, and believe that the expressions are correct (there doesn't seem to be a missing factor of m). Another note (in addition to the symmetry observed in the comment by Paseman) is the (trivial) identity $h_{2m}=g_m+h_m$, leading to another form for the desired bound:\n$$g_{2n+1}(x)+g_{n+1}(x)-h_{2n+1}(x)\\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{n+1}(x).$$</p>\n" }, { "answer_id": 259508, "score": 3, "text": "<p>This is only a comment to @DimaPasechnik, but I cannot put the picture in a comment. The surface to the right of the $x=y$ is a plot of Dima's function (barring mistakes); clearly not convex.</p>\n\n<p><a href=\"https://i.stack.imgur.com/7DSRU.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/7DSRU.jpg\" alt=\"enter image description here\"></a></p>\n" }, { "answer_id": 259489, "score": 2, "text": "<p>I conjecture a stronger statement: that the function $F(x,y)$ of two variables $x,y$ obtained from $F_n(x)$ by substituting $y$ for $x^{2n}$ is convex for $0&lt;y&lt;x&lt;0$. </p>\n\n<p>EDIT2: this conjecture is wrong as stated. See e.g. the plot in the answer by Yaakov Baruch, or rotate the plot in Sage...</p>\n\n<p>While this reduces to checking positive definiteness of \na matrix of bivariate polynomials of degree about 20, obtained from the Hessian of $F$ by clearing common (positive) denominator, this should be possible to make to work by standard real algebraic geometry tools.\nHere is an (ugly) plot of $F$:</p>\n\n<p><a href=\"https://i.stack.imgur.com/l9j8a.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/l9j8a.jpg\" alt=\"plot on [0,1]x[0,1]\"></a> obtained by exporting the plot from jmol in Sage(math) </p>\n\n<pre><code>sage: var('x y')\nsage: F(x,y)=log((x+y^2)*(1+y)*(1-x*y)/(x*(1+x*y)*(1-x^2*y)))\nsage: plot3d(F,[0,1],[0,1])\n</code></pre>\n\n<p>EDIT: comments say that as $y\\to 0+$, the Hessian becomes indefinite. That is, for my original claim to hold, one need to assume $y&gt;\\epsilon&gt;0$, with $\\epsilon$ possibly depending upon $x$. </p>\n" }, { "answer_id": 259370, "score": 21, "text": "<p>Fix $n$, and let $V(x)$ be the polynomial from the question, for which we want to show that it is\npositive on the open interval $(0,1)$. Set\n\\begin{align*}\na(n) &amp;= 1024n^2 - 4096/3n + 320\\\\\nb(n) &amp;= 12288n^3 - 14336n^2 + 4864/3n + 256\\\\\nW(x) &amp;= x^4\\left(V(x)-(n+1)^2(2n+1)^2x^{24n-1}(1-x)^4(a(n)x + b(n)(1-x))\\right).\n\\end{align*}</p>\n\n<p>For all $n\\ge12$ we claim that $(1-x)^6$ divides $W(x)$, and that all the coefficients of $W(x)/(1-x)^6$ are nonnegative. In particular, $W(x)$ is positive for $x&gt;0$. As $a(n)$ and $b(n)$ is positive for all $n\\ge1$, we see that $V(x)$ is positive for all $0&lt;x&lt;1$, provided that $n\\ge12$. The cases $2\\le n\\le 11$ are easily checked directly, for instance by noting that the coefficients of $(1+x)^{\\deg V}V(1/(1+x))$ are positive.</p>\n\n<p>I'm not sure about the easiest kind to prove the assertion about the coefficients of $W(x)/(1-x)^6$. We compute them via\n\\begin{equation}\n\\frac{W(x)}{(1-x)^6}=W(x)\\sum_{k\\ge0}\\binom{k+5}{5}x^k.\n\\end{equation}</p>\n\n<p>All what follows is assisted/confirmed by the Sage code below.</p>\n\n<p>The exponents of $W(x)$ have the form $2ni+j$ for $0\\le i\\le 12$ and $0\\le j\\le 10$. Let $a_{i,j}$ be the corresponding coefficient. It is a polynomial in $n$.</p>\n\n<p>We have\n\\begin{equation} \n\\frac{W(x)}{(1-x)^6} = \\sum_{i,j}a_{i,j}(n)x^{2ni+j}\\sum_k\\binom{k+5}{5}x^k.\n\\end{equation}</p>\n\n<p>With $r\\ge0$ and $0\\le s\\le 2n-1$, the coefficient of $x^{2nr+s}$ in $W(x)\\frac{1}{(1-x)^6}$\nis the sum of all $a_{i,j}(n)\\cdot\\binom{nr+s-ni-j+y}{5}$ for all pairs $(i,j)$ where $i&lt;r$ and $0\\le j\\le 10$, or $i=r$ and $j\\le s$.</p>\n\n<p>We distinguish the cases $0\\le s\\le 9$ from the cases $s\\ge10$. In these latter cases, we write $s=10+y$. Recall that $s\\le 2n-1$, so $10+y=2n-1-o$ for a nonnegative $o$. Upon replacing $n$ with $(11+y+1)/2$, we get polynomials in $n$ and $o$.</p>\n\n<p>In the former cases, the coefficients are polynomials in $n$, which have positive coefficients upon replacing $n$ with $n+12$. So these coefficients are positive for all $n\\ge12$. </p>\n\n<p>In the latter cases it turns out that in all but one case the coefficients are positive. So for nonnegative $y$ and $o$, the values are nonnegative.</p>\n\n<p>For the single exception we see that if we multiply it with a suitable polynomial, the resulting coefficients are positive.</p>\n\n<p><em>Remark (answering Jason's question):</em> It is a known fact that a polynomial $V$ is positive on $(0,1)$ if and only if it is a nonnegative linear combination of polynomials $x^i(1-x)^j$. The problem is that it may involve terms where $i+j$ is bigger than $\\deg V$. (This doesn't happen here, though.) I used an LP solver to play a little with $V$ and $\\frac{V}{(1-x^2)^4x^{2n-2}}$. From that a pattern showed up which lead to a solution.</p>\n\n<p>By the way, $V$ actually seems to be a nonnegative linear combination of $x^i(1-x)^{\\deg V-i}$. This is equivalent to say that all the coefficients of $(1+x)^{\\deg V}V(1/(1+x))$ are nonnegative. It is not hard to compute explicit expressions for these coefficients, but I don't see an easy arguments why they can't be negative.</p>\n\n<pre><code># Formally compute W\nvar('X N')\nP = (1+X^(4*N-1))*(1+X^(2*N))*(1-X^(2*N+1))\nQ = (1+X^(2*N+1))*(1-X^(2*N+2))\nV = P*Q^2*P.diff(X,2)+(P*Q.diff(X))^2-P^2*Q*Q.diff(X,2)-(P.diff(X)*Q)^2\na = 1024*N^2 - 4096/3*N + 320\nb = 12288*N^3 - 14336*N^2 + 4864/3*N + 256\nW = X^4*(V - (2*N + 1)^2 *(N + 1)^2*X^(24*N-1)*(1-X)^4*(a*X + b*(1-X)))\n\n# Check that (1-X)^6 divides W(X)\nprint all(W.diff(X,i)(X=1).polynomial(QQ) == 0 for i in [0..5])\n\n# Compute the coefficients a_ij for the exponents 2ni+j of W. Somewhat\n# clumsy, as I don't know how to deal with polynomials where exponents\n# are symbolic expressions.\n#\n# l is the list of summands of W\nl = [z.canonicalize_radical() for z in W.expand().operands()]\nK.&lt;n&gt; = QQ[]\naij = {(i,j):K(0) for i in [0..12] for j in [0..10]}\nfor term in l:\n c = term(X=1) # get coefficient of term\n e = (term.diff(X)/c)(X=1) # get exponent of term\n c = K(c.polynomial(QQ)) # convert c to proper polynomial in n\n i, j = ZZ(e.diff(N))//2, ZZ(e(N=0)) # Clumsy method to compute pairs (i,j)\n aij[i,j] += c\n\n# Check if coefficients aij[i,j] were correctly computed\nWnew = sum(c(n=N)*X^(2*N*i+j) for (i,j),c in aij.items())\nprint (W-Wnew).canonicalize_radical().is_trivial_zero()\n\ndef bino(k): # binomial coefficient binom(-6,k)=binom(k+5,5)\n return prod(k+5-z for z in range(5))/120\n\n# compute the coefficients of W(X)/(1-X)^6\nK = K.extend_variables(('y', 'o'))\nK.inject_variables()\nfor r in [0..12]:\n for s in [0..10]:\n d = 1 if s == 10 else 0\n # compute coefficient of X^(2nr+s+d*y)\n f = sum(aij[i,j]*bino(2*n*(r-i)+s+d*y-j) for i in [0..r]\n for j in [0..s if i == r else 10])\n # check non-negativity of the polynomial f\n if d == 0:\n # Checks if the coefficient of X^(2nr+s), which is a\n # polynomial in n, has nonnegative coefficients upon replacing\n # n with n+12\n f = f(n=n+12)\n if min(f.coefficients()+[0]) &lt; 0:\n print \"False\"\n else:\n # The coefficient of X^(2nr+10+y), which is a polynomial in n\n # and y, has to be nonnegative for 0 &lt;= 10+y &lt;= 2n-1, so 10+y\n # = 2n-1-o for non-negative o. So upon replacing n with\n # (11+y+o)/2, we get a polynomial in y and o which has to be\n # nonnegative for all nonnegative y and o. In all but one\n # case, this holds because the coeffcients are non-negative.\n f = f(n=(11+y+o)/2)\n if min(f.coefficients()+[0]) &lt; 0:\n c = o^2 + 23*o*y + 1360*y^2 + 99*o + 340*y + 1675\n print min(c.coefficients()+(c*f).coefficients()) &gt;= 0\n</code></pre>\n" }, { "answer_id": 259280, "score": 11, "text": "<p>This proof is mainly based on some of the suggestions in the early version of this answer. Also, we are building here on the comment by Peter Mueller concerning the polynomial $a$ defined below. It turns out that similar techniques can be used for all other nontrivial polynomials arising in the proof, which allowed us to replace multiple uses of Mathematica command Reduce[] in the last version of this answer by showing that the coefficients of the mentioned relevant polynomials are all nonnegative. I have still retained some use of Reduce[] -- but only to quickly make the routine check for $n\\le11$. </p>\n\n<p>Note that \n\\begin{equation}\nG:=\nF_n''(x)\tx^{2-2 n} \\left(x^{2 n}+1\\right)^2 \\left(x^{4 n}+x\\right)^2 \\left(x^{2 n+1}-1\\right)^2\n \\left(x^{2 n+1}+1\\right)^2 \\left(x^{2 n+2}-1\\right)^2\n\\end{equation}\nis a polynomial in $n$, $x$, and $y:=x^{2n}$, of degree $2$ in $n$:\n\\begin{equation}\n\tG=a+bn+cn^2,\n\\end{equation}\nwhere $a,b,c$ are certain polynomials in $x,y$. </p>\n\n<p>The check of $G\\ge0$ for $n=2,\\dots,11$ is straightforward. So, assume $n\\ge12$, whence $0&lt;y&lt;x^{24}&lt;x^{12}$. \nWe also always assume that $0&lt;x&lt;1$, $0&lt;y&lt;x^{24}$ , \n\\begin{equation}\n\tn=\\frac{\\ln y}{2\\ln x}; \\tag{1}\n\\end{equation}\nthe latter relation is of course just another form of the equality $y=x^{2n}$. </p>\n\n<p>Note that \n$0&lt;y&lt;1\\ \\&amp;\\ 0&lt;x&lt;1$ implies $x=1/(1+u)\\ \\&amp;\\ y=1/(1+v)$ for some $u,v\\ge0$. \nAs suggested by Peter Mueller, it turns out that \n\\begin{equation}\n\ta\\; (1 + u)^{10} (1 + v)^{11}|_{x=1/(1 + u), y=1/(1 + v)}\n\\end{equation}\nis a polynomial in $u,v$ with all coefficients nonnegative. \nDetails of all calculations can be seen in the <a href=\"https://www.dropbox.com/s/g87a59uzdsb75m5/SolutionM1.nb?dl=0\" rel=\"nofollow noreferrer\">Mathematica notebook</a> \nand its <a href=\"https://www.dropbox.com/s/k4vlx4brvok099k/SolutionM1.pdf?dl=0\" rel=\"nofollow noreferrer\">pdf image</a></p>\n\n<p>So, \n\\begin{equation}\n\ta\\ge0\n\\end{equation}\nfor all $x,y$ in $(0,1)$. </p>\n\n<p>Next, introduce \n\\begin{equation}\nx_* := \\frac{869}{1000},\\quad \nx_0:=\\frac{7655}{10000},\\quad y_0:= x_0^{24}\\approx0.00164,\\quad\nx_1 := \\frac{985}{1000},\\quad y_1:= x_1^{24}\\approx0.696. \n\\end{equation} </p>\n\n<p>The following cases/subcases/subsubcases are exhaustive: </p>\n\n<p><strong>Case 1: $y\\le y_1$.</strong> Then the condition $0&lt;y\\le y_1\\ \\&amp;\\ 0&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;x^{24}$ implies $x=x_1/(1+u)\\ \\&amp;\\ y=x^{24}/(1+v)$ or $x=x_1+(1-x_1)/(1+u)\\ \\&amp;\\ y=y_1/(1+v)$ for some $u\\ge0$ and $v\\ge0$, depending on whether $x&lt;x_1$ or $x\\ge x_1$. \nIt turns out that \n\\begin{equation}\n\tc\\; (1 + u)^{248} (1 + v)^{10}|_{x=x_1/(1+u)\\ \\&amp;\\ y=x^{24}/(1+v)}\n\\end{equation}\nand \n\\begin{equation}\n\tc\\; (1 + u)^{10} (1 + v)^{10}|_{x=x_1+(1-x_1)/(1+u)\\ \\&amp;\\ y=y_1/(1+v)}\n\\end{equation}\nare each a polynomial in $u,v$ with all coefficients nonnegative. \nSo, \n\\begin{equation}\n\tc\\ge0 \\quad\\text{in Case 1}. \n\\end{equation}\nSo, for the derivative $G'_n$ of $G$ in $n$, we have \n\\begin{equation}\\tag{2}\nG'_n=b+2cn\\ge b+24c=(G'_n)|_{n=12}.\t\n\\end{equation}</p>\n\n<p>Now we have to distinguish two subcases of Case 1, with further subsubcases of Subcase 1.1: </p>\n\n<p><strong>Subcase 1.1: $y\\le y_1$ and ($y\\ge y_0$ or $x\\le x_*$).</strong> </p>\n\n<p><strong>Subsubcase 1.1.1: $y\\le y_1$ and $y\\ge y_0$ and $x\\ge x_1$.</strong> Then $y\\le y_1\\ \\&amp;\\ y\\ge y_0\\ \\&amp;\\ x\\ge x_1 \\ \\&amp;\\ 0&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;x^{24}$ implies \n$$x=x_{111}:=x_1+(1-x_1)/(1+u)\\ \\&amp;\\ y=y_{111}:=y_0+(y_1-y_0)/(1+v)$$\nfor some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=12}\\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=12}\\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p>\n\n<p><strong>Subsubcase 1.1.2: $y\\le y_1$ and $y\\ge y_0$ and $x&lt;x_1$.</strong> Then $x=x_{112}:=x_0+(x_1-x_0)/(1+u)\\ \\&amp;\\ y=y_{112}:=y_0+(x^{24}-y_0)/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112}, y=y_{112}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112},y=y_{112}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p>\n\n<p><strong>Subsubcase 1.1.3: $y\\le y_1$ and $x\\le x_*$.</strong> Then $x=x_{113}:=x_*/(1+u)\\ \\&amp;\\ y=y_{113}:=x^{24}/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=12}\\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p>\n\n<p>So, $G|_{n=12}\\ge0$ and $(G'_n)|_{n=12}\\ge0$ in all the three subsubcases of Subcase 1.1. In view of $(2)$, we have \n$G\\ge0$ in Subcase 1.1. </p>\n\n<p><strong>Subcase 1.2: $y\\le y_1$ and $y&lt;y_0$ and $x&gt;x_*$.</strong> Then, of course, the condition $y\\le y_1$ is redundant. Let here \n\\begin{equation}\n\t\\rho(x):=\\frac7{2(1-x)}. \n\\end{equation}\nThe condition \n$x_*&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;y_0$ implies $x=x_{12}:=x_*+(1-x_*)/(1+u)$ and $y=y_{12}:=y_0/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\t(G'_n)|_{n=\\rho(x)/2}\\; u\\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}}\n\\end{equation}\nand \n\\begin{equation}\n\tG|_{n=\\rho(x)/2}\\; u^2\\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}}\n\\end{equation}\nare each, again, a polynomial in $u,v$ with all coefficients nonnegative. \nSo, similarly to Subcase 1.1, for all $n\\ge\\rho(x)/2$ we get $G\\ge G|_{n=\\rho(x)/2}\\ge0$. \nIn Subcase 1.2, it remains to note that for our particular $n$, as in $(1)$, we indeed have $n\\ge\\rho(x)/2$. This follows because for $y&lt;y_0$ and $x&gt;x_*$ \n\\begin{equation}\n\t\\ln\\frac1y&gt;\\ln\\frac1{y_0}&gt;1.08\\frac7{2(1-x_*)}\\,\\ln\\frac1{x_*}\n\t&gt;\\frac7{2(1-x)}\\,\\ln\\frac1{x}=\\rho(x)\\ln\\frac1{x}. \n\\end{equation}\nSo, $G\\ge0$ in Subcase 1.2 as well. </p>\n\n<p>It remains to consider </p>\n\n<p><strong>Case 2: $y&gt;y_1$.</strong> Then condition $y&gt;y_1\\ \\&amp;\\ 0&lt;x&lt;1\\ \\&amp;\\ 0&lt;y&lt;x^{24}[&lt;x^{12}]$ implies $x=x_2:=x_1+(1-x_1)/(1+u),y=y_2:=y_1+(x^{12}-y_1)/(1+v)$ for some $u\\ge0$ and $v\\ge0$. \nIt turns out that \n\\begin{equation}\n\tb\\;(1 + u)^{140} (1 + v)^{11}|_{x=x_{2},y=y_{2}}\n\\end{equation}\nis also a polynomial in $u,v$ with all coefficients nonnegative. \nHence, $b\\ge0$ in Case 2. \nSo, if $c\\ge0$, then trivially $G\\ge0$ (since $a\\ge0$ always). So, without loss of generality, $c&lt;0$ and hence $G$ is concave in $n$. So, it is enough to bracket the $n$ as in $(1)$ between some simple rational expressions $n_1$ and $n_2$ such that $G|_{n=n_1}\\ge0$ and $G|_{n=n_2}\\ge0$. </p>\n\n<p>In Case 2, $y$ is \"close\" to $1$ and hence, in view of inequality $y&lt;x^{24}$, so is $x$: $x&gt;y_1^{1/12}=x_1=0.985$. So, it is a bit more convenient here to change variables $x, y$ to \"small variables\" $u:=1-x$ and $v:=1-y$. \nThen $0&lt;v&lt;1-y_1\\approx0.304$ and $1-u=x&gt;y^{1/24}=(1-v)^{1/24}&gt;1-v/3$, whence $u&lt;v/3$. Let now $t:=u/v$, so that $0&lt;t&lt;1/3$. \nIt follows, in Case 2, that $t=(1/3)/(1+r)$ and $v=(1-y_1)/(1+s)$ for some $r\\ge0$ and $s\\ge0$. </p>\n\n<p>The mentioned brackets $n_1$ and $n_2$ are respectively $\\ell_1/2$ and $\\ell_2/2$, where<br>\n\\begin{equation}\n\t\\ell_1:=\\frac1t,\\quad\\ell_2:=\\frac{(2 - v) (1 - t v)}{t (1 - v) (2 - t v)}. \n\\end{equation}\nIt is not hard to see (details on this are in the same Mathematica notebook) that then indeed $n_1 &lt; n &lt; n_2$. \nIt remains to show that $G|_{n=n_1}\\ge0$ and $G|_{n=n_2}\\ge0$. </p>\n\n<p>It turns out that \n\\begin{equation}\n\t(G'_n)|_{n=\\ell_1/2;\\,x=1-tv,\\,y=1-v;\\,t=(1/3)/(1+r),\\,v=(1-y_1)/(1+s)}\\; (1 + r)^{10} (1 + s)^{19}\n\\end{equation}\nand \n\\begin{equation}\n\t(G'_n)|_{n=\\ell_2/2;\\,x=1-tv,\\,y=1-v;\\,t=(1/3)/(1+r),\\,v=(1-y_1)/(1+s)}\\; (1 + r)^{10} (1 + s)^{19}\\; q\n\\end{equation}\nare each a polynomial in $r,s$ with all coefficients nonnegative, \nwhere \n\\begin{multline*}\n\tq:=\n\t(1 + r)^{10} (1 + s)^{19} \\\\\n\t\\times(11673186598630578538556565100133681446610566511878526881 \\\\ \n\t+ \n 16777216000000000000000000000000000000000000000000000000 s)^2 \\\\ \n\\times\\big(31853088866210192846185521700044560482203522170626175627 \\\\\n+ \n 33554432000000000000000000000000000000000000000000000000 (r + s + \n r s)\\big)^2. \n\\end{multline*}</p>\n\n<p>The proof is complete. </p>\n" }, { "answer_id": 259260, "score": 10, "text": "<p>I sketch a method that should show $F_n$ is convex for $n$ sufficiently large.</p>\n\n<p>A Taylor expansion gives that $F_n(x) = x^{2n} (1-x)^2 + O(x^{4n-1})$, and so $$F_n''(x) = 2x^{2n-2}[2n^2(1-x)^2 + x^2 - n(1-x)(1+3x)] + O(x^{4n-3}).$$\nThis should show that $F_n''(x) &gt; 0$ for $n$ large, provided $x &lt; 1-\\frac{C}{n}$ for some fixed $C&gt;0$.</p>\n\n<p>On the other hand, according to Mathematica, $F_n''(1) = 4n^2 - \\frac{16}{3} n + \\frac{5}{4}$, which has its largest root at $\\frac{8 + \\sqrt{19}}{12} \\approx 1.02991$. (By the way, there doesn't seem to be any reason that $n$ should be restricted to be an integer, and this is in line with some plots I made indicating $F_n(x)$ is convex for $n &gt; 1.03$.)\nMoreover, Mathematica gives $$\\lim_{n\\rightarrow \\infty} n^{-2}F_n''(1-\\frac{y}{n}) = \\frac{16 e^{4y}}{(1+e^{4y})^2} &gt; 0.$$</p>\n\n<p>Putting these two pieces together should be able to show $F_n(x)$ is convex on $(0,1)$ for $n$ sufficiently large. Unfortunately, it might be very painful to work out all the error terms explicitly.</p>\n" }, { "answer_id": 259207, "score": 6, "text": "<p>Not an answer, just a plot of $F_n(x)$ for $n=2,\\ldots,10$:\n<hr />\n&nbsp; &nbsp; &nbsp; &nbsp; &nbsp;\n<a href=\"https://i.stack.imgur.com/jzWwY.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/jzWwY.jpg\" alt=\"Fn(x)\"></a>\n<hr />\n&nbsp; &nbsp; &nbsp; &nbsp; &nbsp;\n<a href=\"https://i.stack.imgur.com/c8gFL.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/c8gFL.jpg\" alt=\"Fn(x)2\"></a></p>\n\n<hr />\n" }, { "answer_id": 259213, "score": 5, "text": "<p>This is not an answer, just a reformulation into something which looks perhaps tractable (along the lines of the comment by G. Paseman). Define, for $m\\geq 1$ and $x\\in (0,1)$, $$g_m(x):=\\frac{d^2}{dx^2}[\\log(1+x^m)]=\\frac{x^{m-2}(m(m-1)-mx^m)}{(1+x^m)^2}$$ and $$h_m(x):=\\frac{d^2}{dx^2}[\\log(1-x^m)]=-\\frac{x^{m-2}(m(m-1)+mx^m)}{(1-x^m)^2}.$$ Then, for $n\\geq 2$, $$F_n''(x)=g_{4n-1}(x)+g_{2n}(x)+h_{2n+1}(x)-g_{2n+1}(x)-h_{2n+2}(x).$$ </p>\n\n<p>Two observations</p>\n\n<ul>\n<li>The contribution of the first, second, and last terms is non-negative for all $n\\geq 2$ and $x\\in(0,1)$, while that of the remaining terms is non-positive.</li>\n<li>It would therefore suffice to show the following: for $n$ and $x$ as above,\n$$g_{2n+1}(x)-h_{2n+1}(x)\\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{2n+2}(x).$$</li>\n</ul>\n\n<p><strong>Added</strong>: Since I apparently cannot add a comment, I would just note that while it is certainly (always) possible I overlook something, I have just double checked the algebra with maple, and believe that the expressions are correct (there doesn't seem to be a missing factor of m). Another note (in addition to the symmetry observed in the comment by Paseman) is the (trivial) identity $h_{2m}=g_m+h_m$, leading to another form for the desired bound:\n$$g_{2n+1}(x)+g_{n+1}(x)-h_{2n+1}(x)\\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{n+1}(x).$$</p>\n" }, { "answer_id": 259508, "score": 3, "text": "<p>This is only a comment to @DimaPasechnik, but I cannot put the picture in a comment. The surface to the right of the $x=y$ is a plot of Dima's function (barring mistakes); clearly not convex.</p>\n\n<p><a href=\"https://i.stack.imgur.com/7DSRU.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/7DSRU.jpg\" alt=\"enter image description here\"></a></p>\n" }, { "answer_id": 259489, "score": 2, "text": "<p>I conjecture a stronger statement: that the function $F(x,y)$ of two variables $x,y$ obtained from $F_n(x)$ by substituting $y$ for $x^{2n}$ is convex for $0&lt;y&lt;x&lt;0$. </p>\n\n<p>EDIT2: this conjecture is wrong as stated. See e.g. the plot in the answer by Yaakov Baruch, or rotate the plot in Sage...</p>\n\n<p>While this reduces to checking positive definiteness of \na matrix of bivariate polynomials of degree about 20, obtained from the Hessian of $F$ by clearing common (positive) denominator, this should be possible to make to work by standard real algebraic geometry tools.\nHere is an (ugly) plot of $F$:</p>\n\n<p><a href=\"https://i.stack.imgur.com/l9j8a.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/l9j8a.jpg\" alt=\"plot on [0,1]x[0,1]\"></a> obtained by exporting the plot from jmol in Sage(math) </p>\n\n<pre><code>sage: var('x y')\nsage: F(x,y)=log((x+y^2)*(1+y)*(1-x*y)/(x*(1+x*y)*(1-x^2*y)))\nsage: plot3d(F,[0,1],[0,1])\n</code></pre>\n\n<p>EDIT: comments say that as $y\\to 0+$, the Hessian becomes indefinite. That is, for my original claim to hold, one need to assume $y&gt;\\epsilon&gt;0$, with $\\epsilon$ possibly depending upon $x$. </p>\n" } ]
Question: The missing link: an inequality Answer 1: <p>Fix $n$, and let $V(x)$ be the polynomial from the question, for which we want to show that it is positive on the open interval $(0,1)$. Set \begin{align*} a(n) &amp;= 1024n^2 - 4096/3n + 320\\ b(n) &amp;= 12288n^3 - 14336n^2 + 4864/3n + 256\\ W(x) &amp;= x^4\left(V(x)-(n+1)^2(2n+1)^2x^{24n-1}(1-x)^4(a(n)x + b(n)(1-x))\right). \end{align*}</p> <p>For all $n\ge12$ we claim that $(1-x)^6$ divides $W(x)$, and that all the coefficients of $W(x)/(1-x)^6$ are nonnegative. In particular, $W(x)$ is positive for $x&gt;0$. As $a(n)$ and $b(n)$ is positive for all $n\ge1$, we see that $V(x)$ is positive for all $0&lt;x&lt;1$, provided that $n\ge12$. The cases $2\le n\le 11$ are easily checked directly, for instance by noting that the coefficients of $(1+x)^{\deg V}V(1/(1+x))$ are positive.</p> <p>I'm not sure about the easiest kind to prove the assertion about the coefficients of $W(x)/(1-x)^6$. We compute them via \begin{equation} \frac{W(x)}{(1-x)^6}=W(x)\sum_{k\ge0}\binom{k+5}{5}x^k. \end{equation}</p> <p>All what follows is assisted/confirmed by the Sage code below.</p> <p>The exponents of $W(x)$ have the form $2ni+j$ for $0\le i\le 12$ and $0\le j\le 10$. Let $a_{i,j}$ be the corresponding coefficient. It is a polynomial in $n$.</p> <p>We have \begin{equation} \frac{W(x)}{(1-x)^6} = \sum_{i,j}a_{i,j}(n)x^{2ni+j}\sum_k\binom{k+5}{5}x^k. \end{equation}</p> <p>With $r\ge0$ and $0\le s\le 2n-1$, the coefficient of $x^{2nr+s}$ in $W(x)\frac{1}{(1-x)^6}$ is the sum of all $a_{i,j}(n)\cdot\binom{nr+s-ni-j+y}{5}$ for all pairs $(i,j)$ where $i&lt;r$ and $0\le j\le 10$, or $i=r$ and $j\le s$.</p> <p>We distinguish the cases $0\le s\le 9$ from the cases $s\ge10$. In these latter cases, we write $s=10+y$. Recall that $s\le 2n-1$, so $10+y=2n-1-o$ for a nonnegative $o$. Upon replacing $n$ with $(11+y+1)/2$, we get polynomials in $n$ and $o$.</p> <p>In the former cases, the coefficients are polynomials in $n$, which have positive coefficients upon replacing $n$ with $n+12$. So these coefficients are positive for all $n\ge12$. </p> <p>In the latter cases it turns out that in all but one case the coefficients are positive. So for nonnegative $y$ and $o$, the values are nonnegative.</p> <p>For the single exception we see that if we multiply it with a suitable polynomial, the resulting coefficients are positive.</p> <p><em>Remark (answering Jason's question):</em> It is a known fact that a polynomial $V$ is positive on $(0,1)$ if and only if it is a nonnegative linear combination of polynomials $x^i(1-x)^j$. The problem is that it may involve terms where $i+j$ is bigger than $\deg V$. (This doesn't happen here, though.) I used an LP solver to play a little with $V$ and $\frac{V}{(1-x^2)^4x^{2n-2}}$. From that a pattern showed up which lead to a solution.</p> <p>By the way, $V$ actually seems to be a nonnegative linear combination of $x^i(1-x)^{\deg V-i}$. This is equivalent to say that all the coefficients of $(1+x)^{\deg V}V(1/(1+x))$ are nonnegative. It is not hard to compute explicit expressions for these coefficients, but I don't see an easy arguments why they can't be negative.</p> <pre><code># Formally compute W var('X N') P = (1+X^(4*N-1))*(1+X^(2*N))*(1-X^(2*N+1)) Q = (1+X^(2*N+1))*(1-X^(2*N+2)) V = P*Q^2*P.diff(X,2)+(P*Q.diff(X))^2-P^2*Q*Q.diff(X,2)-(P.diff(X)*Q)^2 a = 1024*N^2 - 4096/3*N + 320 b = 12288*N^3 - 14336*N^2 + 4864/3*N + 256 W = X^4*(V - (2*N + 1)^2 *(N + 1)^2*X^(24*N-1)*(1-X)^4*(a*X + b*(1-X))) # Check that (1-X)^6 divides W(X) print all(W.diff(X,i)(X=1).polynomial(QQ) == 0 for i in [0..5]) # Compute the coefficients a_ij for the exponents 2ni+j of W. Somewhat # clumsy, as I don't know how to deal with polynomials where exponents # are symbolic expressions. # # l is the list of summands of W l = [z.canonicalize_radical() for z in W.expand().operands()] K.&lt;n&gt; = QQ[] aij = {(i,j):K(0) for i in [0..12] for j in [0..10]} for term in l: c = term(X=1) # get coefficient of term e = (term.diff(X)/c)(X=1) # get exponent of term c = K(c.polynomial(QQ)) # convert c to proper polynomial in n i, j = ZZ(e.diff(N))//2, ZZ(e(N=0)) # Clumsy method to compute pairs (i,j) aij[i,j] += c # Check if coefficients aij[i,j] were correctly computed Wnew = sum(c(n=N)*X^(2*N*i+j) for (i,j),c in aij.items()) print (W-Wnew).canonicalize_radical().is_trivial_zero() def bino(k): # binomial coefficient binom(-6,k)=binom(k+5,5) return prod(k+5-z for z in range(5))/120 # compute the coefficients of W(X)/(1-X)^6 K = K.extend_variables(('y', 'o')) K.inject_variables() for r in [0..12]: for s in [0..10]: d = 1 if s == 10 else 0 # compute coefficient of X^(2nr+s+d*y) f = sum(aij[i,j]*bino(2*n*(r-i)+s+d*y-j) for i in [0..r] for j in [0..s if i == r else 10]) # check non-negativity of the polynomial f if d == 0: # Checks if the coefficient of X^(2nr+s), which is a # polynomial in n, has nonnegative coefficients upon replacing # n with n+12 f = f(n=n+12) if min(f.coefficients()+[0]) &lt; 0: print "False" else: # The coefficient of X^(2nr+10+y), which is a polynomial in n # and y, has to be nonnegative for 0 &lt;= 10+y &lt;= 2n-1, so 10+y # = 2n-1-o for non-negative o. So upon replacing n with # (11+y+o)/2, we get a polynomial in y and o which has to be # nonnegative for all nonnegative y and o. In all but one # case, this holds because the coeffcients are non-negative. f = f(n=(11+y+o)/2) if min(f.coefficients()+[0]) &lt; 0: c = o^2 + 23*o*y + 1360*y^2 + 99*o + 340*y + 1675 print min(c.coefficients()+(c*f).coefficients()) &gt;= 0 </code></pre> Answer 2: <p>This proof is mainly based on some of the suggestions in the early version of this answer. Also, we are building here on the comment by Peter Mueller concerning the polynomial $a$ defined below. It turns out that similar techniques can be used for all other nontrivial polynomials arising in the proof, which allowed us to replace multiple uses of Mathematica command Reduce[] in the last version of this answer by showing that the coefficients of the mentioned relevant polynomials are all nonnegative. I have still retained some use of Reduce[] -- but only to quickly make the routine check for $n\le11$. </p> <p>Note that \begin{equation} G:= F_n''(x) x^{2-2 n} \left(x^{2 n}+1\right)^2 \left(x^{4 n}+x\right)^2 \left(x^{2 n+1}-1\right)^2 \left(x^{2 n+1}+1\right)^2 \left(x^{2 n+2}-1\right)^2 \end{equation} is a polynomial in $n$, $x$, and $y:=x^{2n}$, of degree $2$ in $n$: \begin{equation} G=a+bn+cn^2, \end{equation} where $a,b,c$ are certain polynomials in $x,y$. </p> <p>The check of $G\ge0$ for $n=2,\dots,11$ is straightforward. So, assume $n\ge12$, whence $0&lt;y&lt;x^{24}&lt;x^{12}$. We also always assume that $0&lt;x&lt;1$, $0&lt;y&lt;x^{24}$ , \begin{equation} n=\frac{\ln y}{2\ln x}; \tag{1} \end{equation} the latter relation is of course just another form of the equality $y=x^{2n}$. </p> <p>Note that $0&lt;y&lt;1\ \&amp;\ 0&lt;x&lt;1$ implies $x=1/(1+u)\ \&amp;\ y=1/(1+v)$ for some $u,v\ge0$. As suggested by Peter Mueller, it turns out that \begin{equation} a\; (1 + u)^{10} (1 + v)^{11}|_{x=1/(1 + u), y=1/(1 + v)} \end{equation} is a polynomial in $u,v$ with all coefficients nonnegative. Details of all calculations can be seen in the <a href="https://www.dropbox.com/s/g87a59uzdsb75m5/SolutionM1.nb?dl=0" rel="nofollow noreferrer">Mathematica notebook</a> and its <a href="https://www.dropbox.com/s/k4vlx4brvok099k/SolutionM1.pdf?dl=0" rel="nofollow noreferrer">pdf image</a></p> <p>So, \begin{equation} a\ge0 \end{equation} for all $x,y$ in $(0,1)$. </p> <p>Next, introduce \begin{equation} x_* := \frac{869}{1000},\quad x_0:=\frac{7655}{10000},\quad y_0:= x_0^{24}\approx0.00164,\quad x_1 := \frac{985}{1000},\quad y_1:= x_1^{24}\approx0.696. \end{equation} </p> <p>The following cases/subcases/subsubcases are exhaustive: </p> <p><strong>Case 1: $y\le y_1$.</strong> Then the condition $0&lt;y\le y_1\ \&amp;\ 0&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;x^{24}$ implies $x=x_1/(1+u)\ \&amp;\ y=x^{24}/(1+v)$ or $x=x_1+(1-x_1)/(1+u)\ \&amp;\ y=y_1/(1+v)$ for some $u\ge0$ and $v\ge0$, depending on whether $x&lt;x_1$ or $x\ge x_1$. It turns out that \begin{equation} c\; (1 + u)^{248} (1 + v)^{10}|_{x=x_1/(1+u)\ \&amp;\ y=x^{24}/(1+v)} \end{equation} and \begin{equation} c\; (1 + u)^{10} (1 + v)^{10}|_{x=x_1+(1-x_1)/(1+u)\ \&amp;\ y=y_1/(1+v)} \end{equation} are each a polynomial in $u,v$ with all coefficients nonnegative. So, \begin{equation} c\ge0 \quad\text{in Case 1}. \end{equation} So, for the derivative $G'_n$ of $G$ in $n$, we have \begin{equation}\tag{2} G'_n=b+2cn\ge b+24c=(G'_n)|_{n=12}. \end{equation}</p> <p>Now we have to distinguish two subcases of Case 1, with further subsubcases of Subcase 1.1: </p> <p><strong>Subcase 1.1: $y\le y_1$ and ($y\ge y_0$ or $x\le x_*$).</strong> </p> <p><strong>Subsubcase 1.1.1: $y\le y_1$ and $y\ge y_0$ and $x\ge x_1$.</strong> Then $y\le y_1\ \&amp;\ y\ge y_0\ \&amp;\ x\ge x_1 \ \&amp;\ 0&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;x^{24}$ implies $$x=x_{111}:=x_1+(1-x_1)/(1+u)\ \&amp;\ y=y_{111}:=y_0+(y_1-y_0)/(1+v)$$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=12}\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}} \end{equation} and \begin{equation} G|_{n=12}\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p> <p><strong>Subsubcase 1.1.2: $y\le y_1$ and $y\ge y_0$ and $x&lt;x_1$.</strong> Then $x=x_{112}:=x_0+(x_1-x_0)/(1+u)\ \&amp;\ y=y_{112}:=y_0+(x^{24}-y_0)/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112}, y=y_{112}} \end{equation} and \begin{equation} G|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112},y=y_{112}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p> <p><strong>Subsubcase 1.1.3: $y\le y_1$ and $x\le x_*$.</strong> Then $x=x_{113}:=x_*/(1+u)\ \&amp;\ y=y_{113}:=x^{24}/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}} \end{equation} and \begin{equation} G|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p> <p>So, $G|_{n=12}\ge0$ and $(G'_n)|_{n=12}\ge0$ in all the three subsubcases of Subcase 1.1. In view of $(2)$, we have $G\ge0$ in Subcase 1.1. </p> <p><strong>Subcase 1.2: $y\le y_1$ and $y&lt;y_0$ and $x&gt;x_*$.</strong> Then, of course, the condition $y\le y_1$ is redundant. Let here \begin{equation} \rho(x):=\frac7{2(1-x)}. \end{equation} The condition $x_*&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;y_0$ implies $x=x_{12}:=x_*+(1-x_*)/(1+u)$ and $y=y_{12}:=y_0/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=\rho(x)/2}\; u\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}} \end{equation} and \begin{equation} G|_{n=\rho(x)/2}\; u^2\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. So, similarly to Subcase 1.1, for all $n\ge\rho(x)/2$ we get $G\ge G|_{n=\rho(x)/2}\ge0$. In Subcase 1.2, it remains to note that for our particular $n$, as in $(1)$, we indeed have $n\ge\rho(x)/2$. This follows because for $y&lt;y_0$ and $x&gt;x_*$ \begin{equation} \ln\frac1y&gt;\ln\frac1{y_0}&gt;1.08\frac7{2(1-x_*)}\,\ln\frac1{x_*} &gt;\frac7{2(1-x)}\,\ln\frac1{x}=\rho(x)\ln\frac1{x}. \end{equation} So, $G\ge0$ in Subcase 1.2 as well. </p> <p>It remains to consider </p> <p><strong>Case 2: $y&gt;y_1$.</strong> Then condition $y&gt;y_1\ \&amp;\ 0&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;x^{24}[&lt;x^{12}]$ implies $x=x_2:=x_1+(1-x_1)/(1+u),y=y_2:=y_1+(x^{12}-y_1)/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} b\;(1 + u)^{140} (1 + v)^{11}|_{x=x_{2},y=y_{2}} \end{equation} is also a polynomial in $u,v$ with all coefficients nonnegative. Hence, $b\ge0$ in Case 2. So, if $c\ge0$, then trivially $G\ge0$ (since $a\ge0$ always). So, without loss of generality, $c&lt;0$ and hence $G$ is concave in $n$. So, it is enough to bracket the $n$ as in $(1)$ between some simple rational expressions $n_1$ and $n_2$ such that $G|_{n=n_1}\ge0$ and $G|_{n=n_2}\ge0$. </p> <p>In Case 2, $y$ is "close" to $1$ and hence, in view of inequality $y&lt;x^{24}$, so is $x$: $x&gt;y_1^{1/12}=x_1=0.985$. So, it is a bit more convenient here to change variables $x, y$ to "small variables" $u:=1-x$ and $v:=1-y$. Then $0&lt;v&lt;1-y_1\approx0.304$ and $1-u=x&gt;y^{1/24}=(1-v)^{1/24}&gt;1-v/3$, whence $u&lt;v/3$. Let now $t:=u/v$, so that $0&lt;t&lt;1/3$. It follows, in Case 2, that $t=(1/3)/(1+r)$ and $v=(1-y_1)/(1+s)$ for some $r\ge0$ and $s\ge0$. </p> <p>The mentioned brackets $n_1$ and $n_2$ are respectively $\ell_1/2$ and $\ell_2/2$, where<br> \begin{equation} \ell_1:=\frac1t,\quad\ell_2:=\frac{(2 - v) (1 - t v)}{t (1 - v) (2 - t v)}. \end{equation} It is not hard to see (details on this are in the same Mathematica notebook) that then indeed $n_1 &lt; n &lt; n_2$. It remains to show that $G|_{n=n_1}\ge0$ and $G|_{n=n_2}\ge0$. </p> <p>It turns out that \begin{equation} (G'_n)|_{n=\ell_1/2;\,x=1-tv,\,y=1-v;\,t=(1/3)/(1+r),\,v=(1-y_1)/(1+s)}\; (1 + r)^{10} (1 + s)^{19} \end{equation} and \begin{equation} (G'_n)|_{n=\ell_2/2;\,x=1-tv,\,y=1-v;\,t=(1/3)/(1+r),\,v=(1-y_1)/(1+s)}\; (1 + r)^{10} (1 + s)^{19}\; q \end{equation} are each a polynomial in $r,s$ with all coefficients nonnegative, where \begin{multline*} q:= (1 + r)^{10} (1 + s)^{19} \\ \times(11673186598630578538556565100133681446610566511878526881 \\ + 16777216000000000000000000000000000000000000000000000000 s)^2 \\ \times\big(31853088866210192846185521700044560482203522170626175627 \\ + 33554432000000000000000000000000000000000000000000000000 (r + s + r s)\big)^2. \end{multline*}</p> <p>The proof is complete. </p> Answer 3: <p>I sketch a method that should show $F_n$ is convex for $n$ sufficiently large.</p> <p>A Taylor expansion gives that $F_n(x) = x^{2n} (1-x)^2 + O(x^{4n-1})$, and so $$F_n''(x) = 2x^{2n-2}[2n^2(1-x)^2 + x^2 - n(1-x)(1+3x)] + O(x^{4n-3}).$$ This should show that $F_n''(x) &gt; 0$ for $n$ large, provided $x &lt; 1-\frac{C}{n}$ for some fixed $C&gt;0$.</p> <p>On the other hand, according to Mathematica, $F_n''(1) = 4n^2 - \frac{16}{3} n + \frac{5}{4}$, which has its largest root at $\frac{8 + \sqrt{19}}{12} \approx 1.02991$. (By the way, there doesn't seem to be any reason that $n$ should be restricted to be an integer, and this is in line with some plots I made indicating $F_n(x)$ is convex for $n &gt; 1.03$.) Moreover, Mathematica gives $$\lim_{n\rightarrow \infty} n^{-2}F_n''(1-\frac{y}{n}) = \frac{16 e^{4y}}{(1+e^{4y})^2} &gt; 0.$$</p> <p>Putting these two pieces together should be able to show $F_n(x)$ is convex on $(0,1)$ for $n$ sufficiently large. Unfortunately, it might be very painful to work out all the error terms explicitly.</p> Answer 4: <p>Not an answer, just a plot of $F_n(x)$ for $n=2,\ldots,10$: <hr /> &nbsp; &nbsp; &nbsp; &nbsp; &nbsp; <a href="https://i.stack.imgur.com/jzWwY.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/jzWwY.jpg" alt="Fn(x)"></a> <hr /> &nbsp; &nbsp; &nbsp; &nbsp; &nbsp; <a href="https://i.stack.imgur.com/c8gFL.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/c8gFL.jpg" alt="Fn(x)2"></a></p> <hr /> Answer 5: <p>This is not an answer, just a reformulation into something which looks perhaps tractable (along the lines of the comment by G. Paseman). Define, for $m\geq 1$ and $x\in (0,1)$, $$g_m(x):=\frac{d^2}{dx^2}[\log(1+x^m)]=\frac{x^{m-2}(m(m-1)-mx^m)}{(1+x^m)^2}$$ and $$h_m(x):=\frac{d^2}{dx^2}[\log(1-x^m)]=-\frac{x^{m-2}(m(m-1)+mx^m)}{(1-x^m)^2}.$$ Then, for $n\geq 2$, $$F_n''(x)=g_{4n-1}(x)+g_{2n}(x)+h_{2n+1}(x)-g_{2n+1}(x)-h_{2n+2}(x).$$ </p> <p>Two observations</p> <ul> <li>The contribution of the first, second, and last terms is non-negative for all $n\geq 2$ and $x\in(0,1)$, while that of the remaining terms is non-positive.</li> <li>It would therefore suffice to show the following: for $n$ and $x$ as above, $$g_{2n+1}(x)-h_{2n+1}(x)\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{2n+2}(x).$$</li> </ul> <p><strong>Added</strong>: Since I apparently cannot add a comment, I would just note that while it is certainly (always) possible I overlook something, I have just double checked the algebra with maple, and believe that the expressions are correct (there doesn't seem to be a missing factor of m). Another note (in addition to the symmetry observed in the comment by Paseman) is the (trivial) identity $h_{2m}=g_m+h_m$, leading to another form for the desired bound: $$g_{2n+1}(x)+g_{n+1}(x)-h_{2n+1}(x)\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{n+1}(x).$$</p> Answer 6: <p>This is only a comment to @DimaPasechnik, but I cannot put the picture in a comment. The surface to the right of the $x=y$ is a plot of Dima's function (barring mistakes); clearly not convex.</p> <p><a href="https://i.stack.imgur.com/7DSRU.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/7DSRU.jpg" alt="enter image description here"></a></p> Answer 7: <p>I conjecture a stronger statement: that the function $F(x,y)$ of two variables $x,y$ obtained from $F_n(x)$ by substituting $y$ for $x^{2n}$ is convex for $0&lt;y&lt;x&lt;0$. </p> <p>EDIT2: this conjecture is wrong as stated. See e.g. the plot in the answer by Yaakov Baruch, or rotate the plot in Sage...</p> <p>While this reduces to checking positive definiteness of a matrix of bivariate polynomials of degree about 20, obtained from the Hessian of $F$ by clearing common (positive) denominator, this should be possible to make to work by standard real algebraic geometry tools. Here is an (ugly) plot of $F$:</p> <p><a href="https://i.stack.imgur.com/l9j8a.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/l9j8a.jpg" alt="plot on [0,1]x[0,1]"></a> obtained by exporting the plot from jmol in Sage(math) </p> <pre><code>sage: var('x y') sage: F(x,y)=log((x+y^2)*(1+y)*(1-x*y)/(x*(1+x*y)*(1-x^2*y))) sage: plot3d(F,[0,1],[0,1]) </code></pre> <p>EDIT: comments say that as $y\to 0+$, the Hessian becomes indefinite. That is, for my original claim to hold, one need to assume $y&gt;\epsilon&gt;0$, with $\epsilon$ possibly depending upon $x$. </p> Answer 8: <p>Fix $n$, and let $V(x)$ be the polynomial from the question, for which we want to show that it is positive on the open interval $(0,1)$. Set \begin{align*} a(n) &amp;= 1024n^2 - 4096/3n + 320\\ b(n) &amp;= 12288n^3 - 14336n^2 + 4864/3n + 256\\ W(x) &amp;= x^4\left(V(x)-(n+1)^2(2n+1)^2x^{24n-1}(1-x)^4(a(n)x + b(n)(1-x))\right). \end{align*}</p> <p>For all $n\ge12$ we claim that $(1-x)^6$ divides $W(x)$, and that all the coefficients of $W(x)/(1-x)^6$ are nonnegative. In particular, $W(x)$ is positive for $x&gt;0$. As $a(n)$ and $b(n)$ is positive for all $n\ge1$, we see that $V(x)$ is positive for all $0&lt;x&lt;1$, provided that $n\ge12$. The cases $2\le n\le 11$ are easily checked directly, for instance by noting that the coefficients of $(1+x)^{\deg V}V(1/(1+x))$ are positive.</p> <p>I'm not sure about the easiest kind to prove the assertion about the coefficients of $W(x)/(1-x)^6$. We compute them via \begin{equation} \frac{W(x)}{(1-x)^6}=W(x)\sum_{k\ge0}\binom{k+5}{5}x^k. \end{equation}</p> <p>All what follows is assisted/confirmed by the Sage code below.</p> <p>The exponents of $W(x)$ have the form $2ni+j$ for $0\le i\le 12$ and $0\le j\le 10$. Let $a_{i,j}$ be the corresponding coefficient. It is a polynomial in $n$.</p> <p>We have \begin{equation} \frac{W(x)}{(1-x)^6} = \sum_{i,j}a_{i,j}(n)x^{2ni+j}\sum_k\binom{k+5}{5}x^k. \end{equation}</p> <p>With $r\ge0$ and $0\le s\le 2n-1$, the coefficient of $x^{2nr+s}$ in $W(x)\frac{1}{(1-x)^6}$ is the sum of all $a_{i,j}(n)\cdot\binom{nr+s-ni-j+y}{5}$ for all pairs $(i,j)$ where $i&lt;r$ and $0\le j\le 10$, or $i=r$ and $j\le s$.</p> <p>We distinguish the cases $0\le s\le 9$ from the cases $s\ge10$. In these latter cases, we write $s=10+y$. Recall that $s\le 2n-1$, so $10+y=2n-1-o$ for a nonnegative $o$. Upon replacing $n$ with $(11+y+1)/2$, we get polynomials in $n$ and $o$.</p> <p>In the former cases, the coefficients are polynomials in $n$, which have positive coefficients upon replacing $n$ with $n+12$. So these coefficients are positive for all $n\ge12$. </p> <p>In the latter cases it turns out that in all but one case the coefficients are positive. So for nonnegative $y$ and $o$, the values are nonnegative.</p> <p>For the single exception we see that if we multiply it with a suitable polynomial, the resulting coefficients are positive.</p> <p><em>Remark (answering Jason's question):</em> It is a known fact that a polynomial $V$ is positive on $(0,1)$ if and only if it is a nonnegative linear combination of polynomials $x^i(1-x)^j$. The problem is that it may involve terms where $i+j$ is bigger than $\deg V$. (This doesn't happen here, though.) I used an LP solver to play a little with $V$ and $\frac{V}{(1-x^2)^4x^{2n-2}}$. From that a pattern showed up which lead to a solution.</p> <p>By the way, $V$ actually seems to be a nonnegative linear combination of $x^i(1-x)^{\deg V-i}$. This is equivalent to say that all the coefficients of $(1+x)^{\deg V}V(1/(1+x))$ are nonnegative. It is not hard to compute explicit expressions for these coefficients, but I don't see an easy arguments why they can't be negative.</p> <pre><code># Formally compute W var('X N') P = (1+X^(4*N-1))*(1+X^(2*N))*(1-X^(2*N+1)) Q = (1+X^(2*N+1))*(1-X^(2*N+2)) V = P*Q^2*P.diff(X,2)+(P*Q.diff(X))^2-P^2*Q*Q.diff(X,2)-(P.diff(X)*Q)^2 a = 1024*N^2 - 4096/3*N + 320 b = 12288*N^3 - 14336*N^2 + 4864/3*N + 256 W = X^4*(V - (2*N + 1)^2 *(N + 1)^2*X^(24*N-1)*(1-X)^4*(a*X + b*(1-X))) # Check that (1-X)^6 divides W(X) print all(W.diff(X,i)(X=1).polynomial(QQ) == 0 for i in [0..5]) # Compute the coefficients a_ij for the exponents 2ni+j of W. Somewhat # clumsy, as I don't know how to deal with polynomials where exponents # are symbolic expressions. # # l is the list of summands of W l = [z.canonicalize_radical() for z in W.expand().operands()] K.&lt;n&gt; = QQ[] aij = {(i,j):K(0) for i in [0..12] for j in [0..10]} for term in l: c = term(X=1) # get coefficient of term e = (term.diff(X)/c)(X=1) # get exponent of term c = K(c.polynomial(QQ)) # convert c to proper polynomial in n i, j = ZZ(e.diff(N))//2, ZZ(e(N=0)) # Clumsy method to compute pairs (i,j) aij[i,j] += c # Check if coefficients aij[i,j] were correctly computed Wnew = sum(c(n=N)*X^(2*N*i+j) for (i,j),c in aij.items()) print (W-Wnew).canonicalize_radical().is_trivial_zero() def bino(k): # binomial coefficient binom(-6,k)=binom(k+5,5) return prod(k+5-z for z in range(5))/120 # compute the coefficients of W(X)/(1-X)^6 K = K.extend_variables(('y', 'o')) K.inject_variables() for r in [0..12]: for s in [0..10]: d = 1 if s == 10 else 0 # compute coefficient of X^(2nr+s+d*y) f = sum(aij[i,j]*bino(2*n*(r-i)+s+d*y-j) for i in [0..r] for j in [0..s if i == r else 10]) # check non-negativity of the polynomial f if d == 0: # Checks if the coefficient of X^(2nr+s), which is a # polynomial in n, has nonnegative coefficients upon replacing # n with n+12 f = f(n=n+12) if min(f.coefficients()+[0]) &lt; 0: print "False" else: # The coefficient of X^(2nr+10+y), which is a polynomial in n # and y, has to be nonnegative for 0 &lt;= 10+y &lt;= 2n-1, so 10+y # = 2n-1-o for non-negative o. So upon replacing n with # (11+y+o)/2, we get a polynomial in y and o which has to be # nonnegative for all nonnegative y and o. In all but one # case, this holds because the coeffcients are non-negative. f = f(n=(11+y+o)/2) if min(f.coefficients()+[0]) &lt; 0: c = o^2 + 23*o*y + 1360*y^2 + 99*o + 340*y + 1675 print min(c.coefficients()+(c*f).coefficients()) &gt;= 0 </code></pre> Answer 9: <p>This proof is mainly based on some of the suggestions in the early version of this answer. Also, we are building here on the comment by Peter Mueller concerning the polynomial $a$ defined below. It turns out that similar techniques can be used for all other nontrivial polynomials arising in the proof, which allowed us to replace multiple uses of Mathematica command Reduce[] in the last version of this answer by showing that the coefficients of the mentioned relevant polynomials are all nonnegative. I have still retained some use of Reduce[] -- but only to quickly make the routine check for $n\le11$. </p> <p>Note that \begin{equation} G:= F_n''(x) x^{2-2 n} \left(x^{2 n}+1\right)^2 \left(x^{4 n}+x\right)^2 \left(x^{2 n+1}-1\right)^2 \left(x^{2 n+1}+1\right)^2 \left(x^{2 n+2}-1\right)^2 \end{equation} is a polynomial in $n$, $x$, and $y:=x^{2n}$, of degree $2$ in $n$: \begin{equation} G=a+bn+cn^2, \end{equation} where $a,b,c$ are certain polynomials in $x,y$. </p> <p>The check of $G\ge0$ for $n=2,\dots,11$ is straightforward. So, assume $n\ge12$, whence $0&lt;y&lt;x^{24}&lt;x^{12}$. We also always assume that $0&lt;x&lt;1$, $0&lt;y&lt;x^{24}$ , \begin{equation} n=\frac{\ln y}{2\ln x}; \tag{1} \end{equation} the latter relation is of course just another form of the equality $y=x^{2n}$. </p> <p>Note that $0&lt;y&lt;1\ \&amp;\ 0&lt;x&lt;1$ implies $x=1/(1+u)\ \&amp;\ y=1/(1+v)$ for some $u,v\ge0$. As suggested by Peter Mueller, it turns out that \begin{equation} a\; (1 + u)^{10} (1 + v)^{11}|_{x=1/(1 + u), y=1/(1 + v)} \end{equation} is a polynomial in $u,v$ with all coefficients nonnegative. Details of all calculations can be seen in the <a href="https://www.dropbox.com/s/g87a59uzdsb75m5/SolutionM1.nb?dl=0" rel="nofollow noreferrer">Mathematica notebook</a> and its <a href="https://www.dropbox.com/s/k4vlx4brvok099k/SolutionM1.pdf?dl=0" rel="nofollow noreferrer">pdf image</a></p> <p>So, \begin{equation} a\ge0 \end{equation} for all $x,y$ in $(0,1)$. </p> <p>Next, introduce \begin{equation} x_* := \frac{869}{1000},\quad x_0:=\frac{7655}{10000},\quad y_0:= x_0^{24}\approx0.00164,\quad x_1 := \frac{985}{1000},\quad y_1:= x_1^{24}\approx0.696. \end{equation} </p> <p>The following cases/subcases/subsubcases are exhaustive: </p> <p><strong>Case 1: $y\le y_1$.</strong> Then the condition $0&lt;y\le y_1\ \&amp;\ 0&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;x^{24}$ implies $x=x_1/(1+u)\ \&amp;\ y=x^{24}/(1+v)$ or $x=x_1+(1-x_1)/(1+u)\ \&amp;\ y=y_1/(1+v)$ for some $u\ge0$ and $v\ge0$, depending on whether $x&lt;x_1$ or $x\ge x_1$. It turns out that \begin{equation} c\; (1 + u)^{248} (1 + v)^{10}|_{x=x_1/(1+u)\ \&amp;\ y=x^{24}/(1+v)} \end{equation} and \begin{equation} c\; (1 + u)^{10} (1 + v)^{10}|_{x=x_1+(1-x_1)/(1+u)\ \&amp;\ y=y_1/(1+v)} \end{equation} are each a polynomial in $u,v$ with all coefficients nonnegative. So, \begin{equation} c\ge0 \quad\text{in Case 1}. \end{equation} So, for the derivative $G'_n$ of $G$ in $n$, we have \begin{equation}\tag{2} G'_n=b+2cn\ge b+24c=(G'_n)|_{n=12}. \end{equation}</p> <p>Now we have to distinguish two subcases of Case 1, with further subsubcases of Subcase 1.1: </p> <p><strong>Subcase 1.1: $y\le y_1$ and ($y\ge y_0$ or $x\le x_*$).</strong> </p> <p><strong>Subsubcase 1.1.1: $y\le y_1$ and $y\ge y_0$ and $x\ge x_1$.</strong> Then $y\le y_1\ \&amp;\ y\ge y_0\ \&amp;\ x\ge x_1 \ \&amp;\ 0&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;x^{24}$ implies $$x=x_{111}:=x_1+(1-x_1)/(1+u)\ \&amp;\ y=y_{111}:=y_0+(y_1-y_0)/(1+v)$$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=12}\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}} \end{equation} and \begin{equation} G|_{n=12}\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p> <p><strong>Subsubcase 1.1.2: $y\le y_1$ and $y\ge y_0$ and $x&lt;x_1$.</strong> Then $x=x_{112}:=x_0+(x_1-x_0)/(1+u)\ \&amp;\ y=y_{112}:=y_0+(x^{24}-y_0)/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112}, y=y_{112}} \end{equation} and \begin{equation} G|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112},y=y_{112}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p> <p><strong>Subsubcase 1.1.3: $y\le y_1$ and $x\le x_*$.</strong> Then $x=x_{113}:=x_*/(1+u)\ \&amp;\ y=y_{113}:=x^{24}/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}} \end{equation} and \begin{equation} G|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p> <p>So, $G|_{n=12}\ge0$ and $(G'_n)|_{n=12}\ge0$ in all the three subsubcases of Subcase 1.1. In view of $(2)$, we have $G\ge0$ in Subcase 1.1. </p> <p><strong>Subcase 1.2: $y\le y_1$ and $y&lt;y_0$ and $x&gt;x_*$.</strong> Then, of course, the condition $y\le y_1$ is redundant. Let here \begin{equation} \rho(x):=\frac7{2(1-x)}. \end{equation} The condition $x_*&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;y_0$ implies $x=x_{12}:=x_*+(1-x_*)/(1+u)$ and $y=y_{12}:=y_0/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=\rho(x)/2}\; u\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}} \end{equation} and \begin{equation} G|_{n=\rho(x)/2}\; u^2\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. So, similarly to Subcase 1.1, for all $n\ge\rho(x)/2$ we get $G\ge G|_{n=\rho(x)/2}\ge0$. In Subcase 1.2, it remains to note that for our particular $n$, as in $(1)$, we indeed have $n\ge\rho(x)/2$. This follows because for $y&lt;y_0$ and $x&gt;x_*$ \begin{equation} \ln\frac1y&gt;\ln\frac1{y_0}&gt;1.08\frac7{2(1-x_*)}\,\ln\frac1{x_*} &gt;\frac7{2(1-x)}\,\ln\frac1{x}=\rho(x)\ln\frac1{x}. \end{equation} So, $G\ge0$ in Subcase 1.2 as well. </p> <p>It remains to consider </p> <p><strong>Case 2: $y&gt;y_1$.</strong> Then condition $y&gt;y_1\ \&amp;\ 0&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;x^{24}[&lt;x^{12}]$ implies $x=x_2:=x_1+(1-x_1)/(1+u),y=y_2:=y_1+(x^{12}-y_1)/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} b\;(1 + u)^{140} (1 + v)^{11}|_{x=x_{2},y=y_{2}} \end{equation} is also a polynomial in $u,v$ with all coefficients nonnegative. Hence, $b\ge0$ in Case 2. So, if $c\ge0$, then trivially $G\ge0$ (since $a\ge0$ always). So, without loss of generality, $c&lt;0$ and hence $G$ is concave in $n$. So, it is enough to bracket the $n$ as in $(1)$ between some simple rational expressions $n_1$ and $n_2$ such that $G|_{n=n_1}\ge0$ and $G|_{n=n_2}\ge0$. </p> <p>In Case 2, $y$ is "close" to $1$ and hence, in view of inequality $y&lt;x^{24}$, so is $x$: $x&gt;y_1^{1/12}=x_1=0.985$. So, it is a bit more convenient here to change variables $x, y$ to "small variables" $u:=1-x$ and $v:=1-y$. Then $0&lt;v&lt;1-y_1\approx0.304$ and $1-u=x&gt;y^{1/24}=(1-v)^{1/24}&gt;1-v/3$, whence $u&lt;v/3$. Let now $t:=u/v$, so that $0&lt;t&lt;1/3$. It follows, in Case 2, that $t=(1/3)/(1+r)$ and $v=(1-y_1)/(1+s)$ for some $r\ge0$ and $s\ge0$. </p> <p>The mentioned brackets $n_1$ and $n_2$ are respectively $\ell_1/2$ and $\ell_2/2$, where<br> \begin{equation} \ell_1:=\frac1t,\quad\ell_2:=\frac{(2 - v) (1 - t v)}{t (1 - v) (2 - t v)}. \end{equation} It is not hard to see (details on this are in the same Mathematica notebook) that then indeed $n_1 &lt; n &lt; n_2$. It remains to show that $G|_{n=n_1}\ge0$ and $G|_{n=n_2}\ge0$. </p> <p>It turns out that \begin{equation} (G'_n)|_{n=\ell_1/2;\,x=1-tv,\,y=1-v;\,t=(1/3)/(1+r),\,v=(1-y_1)/(1+s)}\; (1 + r)^{10} (1 + s)^{19} \end{equation} and \begin{equation} (G'_n)|_{n=\ell_2/2;\,x=1-tv,\,y=1-v;\,t=(1/3)/(1+r),\,v=(1-y_1)/(1+s)}\; (1 + r)^{10} (1 + s)^{19}\; q \end{equation} are each a polynomial in $r,s$ with all coefficients nonnegative, where \begin{multline*} q:= (1 + r)^{10} (1 + s)^{19} \\ \times(11673186598630578538556565100133681446610566511878526881 \\ + 16777216000000000000000000000000000000000000000000000000 s)^2 \\ \times\big(31853088866210192846185521700044560482203522170626175627 \\ + 33554432000000000000000000000000000000000000000000000000 (r + s + r s)\big)^2. \end{multline*}</p> <p>The proof is complete. </p> Answer 10: <p>I sketch a method that should show $F_n$ is convex for $n$ sufficiently large.</p> <p>A Taylor expansion gives that $F_n(x) = x^{2n} (1-x)^2 + O(x^{4n-1})$, and so $$F_n''(x) = 2x^{2n-2}[2n^2(1-x)^2 + x^2 - n(1-x)(1+3x)] + O(x^{4n-3}).$$ This should show that $F_n''(x) &gt; 0$ for $n$ large, provided $x &lt; 1-\frac{C}{n}$ for some fixed $C&gt;0$.</p> <p>On the other hand, according to Mathematica, $F_n''(1) = 4n^2 - \frac{16}{3} n + \frac{5}{4}$, which has its largest root at $\frac{8 + \sqrt{19}}{12} \approx 1.02991$. (By the way, there doesn't seem to be any reason that $n$ should be restricted to be an integer, and this is in line with some plots I made indicating $F_n(x)$ is convex for $n &gt; 1.03$.) Moreover, Mathematica gives $$\lim_{n\rightarrow \infty} n^{-2}F_n''(1-\frac{y}{n}) = \frac{16 e^{4y}}{(1+e^{4y})^2} &gt; 0.$$</p> <p>Putting these two pieces together should be able to show $F_n(x)$ is convex on $(0,1)$ for $n$ sufficiently large. Unfortunately, it might be very painful to work out all the error terms explicitly.</p> Answer 11: <p>Not an answer, just a plot of $F_n(x)$ for $n=2,\ldots,10$: <hr /> &nbsp; &nbsp; &nbsp; &nbsp; &nbsp; <a href="https://i.stack.imgur.com/jzWwY.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/jzWwY.jpg" alt="Fn(x)"></a> <hr /> &nbsp; &nbsp; &nbsp; &nbsp; &nbsp; <a href="https://i.stack.imgur.com/c8gFL.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/c8gFL.jpg" alt="Fn(x)2"></a></p> <hr /> Answer 12: <p>This is not an answer, just a reformulation into something which looks perhaps tractable (along the lines of the comment by G. Paseman). Define, for $m\geq 1$ and $x\in (0,1)$, $$g_m(x):=\frac{d^2}{dx^2}[\log(1+x^m)]=\frac{x^{m-2}(m(m-1)-mx^m)}{(1+x^m)^2}$$ and $$h_m(x):=\frac{d^2}{dx^2}[\log(1-x^m)]=-\frac{x^{m-2}(m(m-1)+mx^m)}{(1-x^m)^2}.$$ Then, for $n\geq 2$, $$F_n''(x)=g_{4n-1}(x)+g_{2n}(x)+h_{2n+1}(x)-g_{2n+1}(x)-h_{2n+2}(x).$$ </p> <p>Two observations</p> <ul> <li>The contribution of the first, second, and last terms is non-negative for all $n\geq 2$ and $x\in(0,1)$, while that of the remaining terms is non-positive.</li> <li>It would therefore suffice to show the following: for $n$ and $x$ as above, $$g_{2n+1}(x)-h_{2n+1}(x)\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{2n+2}(x).$$</li> </ul> <p><strong>Added</strong>: Since I apparently cannot add a comment, I would just note that while it is certainly (always) possible I overlook something, I have just double checked the algebra with maple, and believe that the expressions are correct (there doesn't seem to be a missing factor of m). Another note (in addition to the symmetry observed in the comment by Paseman) is the (trivial) identity $h_{2m}=g_m+h_m$, leading to another form for the desired bound: $$g_{2n+1}(x)+g_{n+1}(x)-h_{2n+1}(x)\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{n+1}(x).$$</p> Answer 13: <p>This is only a comment to @DimaPasechnik, but I cannot put the picture in a comment. The surface to the right of the $x=y$ is a plot of Dima's function (barring mistakes); clearly not convex.</p> <p><a href="https://i.stack.imgur.com/7DSRU.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/7DSRU.jpg" alt="enter image description here"></a></p> Answer 14: <p>I conjecture a stronger statement: that the function $F(x,y)$ of two variables $x,y$ obtained from $F_n(x)$ by substituting $y$ for $x^{2n}$ is convex for $0&lt;y&lt;x&lt;0$. </p> <p>EDIT2: this conjecture is wrong as stated. See e.g. the plot in the answer by Yaakov Baruch, or rotate the plot in Sage...</p> <p>While this reduces to checking positive definiteness of a matrix of bivariate polynomials of degree about 20, obtained from the Hessian of $F$ by clearing common (positive) denominator, this should be possible to make to work by standard real algebraic geometry tools. Here is an (ugly) plot of $F$:</p> <p><a href="https://i.stack.imgur.com/l9j8a.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/l9j8a.jpg" alt="plot on [0,1]x[0,1]"></a> obtained by exporting the plot from jmol in Sage(math) </p> <pre><code>sage: var('x y') sage: F(x,y)=log((x+y^2)*(1+y)*(1-x*y)/(x*(1+x*y)*(1-x^2*y))) sage: plot3d(F,[0,1],[0,1]) </code></pre> <p>EDIT: comments say that as $y\to 0+$, the Hessian becomes indefinite. That is, for my original claim to hold, one need to assume $y&gt;\epsilon&gt;0$, with $\epsilon$ possibly depending upon $x$. </p> Answer 15: <p>Fix $n$, and let $V(x)$ be the polynomial from the question, for which we want to show that it is positive on the open interval $(0,1)$. Set \begin{align*} a(n) &amp;= 1024n^2 - 4096/3n + 320\\ b(n) &amp;= 12288n^3 - 14336n^2 + 4864/3n + 256\\ W(x) &amp;= x^4\left(V(x)-(n+1)^2(2n+1)^2x^{24n-1}(1-x)^4(a(n)x + b(n)(1-x))\right). \end{align*}</p> <p>For all $n\ge12$ we claim that $(1-x)^6$ divides $W(x)$, and that all the coefficients of $W(x)/(1-x)^6$ are nonnegative. In particular, $W(x)$ is positive for $x&gt;0$. As $a(n)$ and $b(n)$ is positive for all $n\ge1$, we see that $V(x)$ is positive for all $0&lt;x&lt;1$, provided that $n\ge12$. The cases $2\le n\le 11$ are easily checked directly, for instance by noting that the coefficients of $(1+x)^{\deg V}V(1/(1+x))$ are positive.</p> <p>I'm not sure about the easiest kind to prove the assertion about the coefficients of $W(x)/(1-x)^6$. We compute them via \begin{equation} \frac{W(x)}{(1-x)^6}=W(x)\sum_{k\ge0}\binom{k+5}{5}x^k. \end{equation}</p> <p>All what follows is assisted/confirmed by the Sage code below.</p> <p>The exponents of $W(x)$ have the form $2ni+j$ for $0\le i\le 12$ and $0\le j\le 10$. Let $a_{i,j}$ be the corresponding coefficient. It is a polynomial in $n$.</p> <p>We have \begin{equation} \frac{W(x)}{(1-x)^6} = \sum_{i,j}a_{i,j}(n)x^{2ni+j}\sum_k\binom{k+5}{5}x^k. \end{equation}</p> <p>With $r\ge0$ and $0\le s\le 2n-1$, the coefficient of $x^{2nr+s}$ in $W(x)\frac{1}{(1-x)^6}$ is the sum of all $a_{i,j}(n)\cdot\binom{nr+s-ni-j+y}{5}$ for all pairs $(i,j)$ where $i&lt;r$ and $0\le j\le 10$, or $i=r$ and $j\le s$.</p> <p>We distinguish the cases $0\le s\le 9$ from the cases $s\ge10$. In these latter cases, we write $s=10+y$. Recall that $s\le 2n-1$, so $10+y=2n-1-o$ for a nonnegative $o$. Upon replacing $n$ with $(11+y+1)/2$, we get polynomials in $n$ and $o$.</p> <p>In the former cases, the coefficients are polynomials in $n$, which have positive coefficients upon replacing $n$ with $n+12$. So these coefficients are positive for all $n\ge12$. </p> <p>In the latter cases it turns out that in all but one case the coefficients are positive. So for nonnegative $y$ and $o$, the values are nonnegative.</p> <p>For the single exception we see that if we multiply it with a suitable polynomial, the resulting coefficients are positive.</p> <p><em>Remark (answering Jason's question):</em> It is a known fact that a polynomial $V$ is positive on $(0,1)$ if and only if it is a nonnegative linear combination of polynomials $x^i(1-x)^j$. The problem is that it may involve terms where $i+j$ is bigger than $\deg V$. (This doesn't happen here, though.) I used an LP solver to play a little with $V$ and $\frac{V}{(1-x^2)^4x^{2n-2}}$. From that a pattern showed up which lead to a solution.</p> <p>By the way, $V$ actually seems to be a nonnegative linear combination of $x^i(1-x)^{\deg V-i}$. This is equivalent to say that all the coefficients of $(1+x)^{\deg V}V(1/(1+x))$ are nonnegative. It is not hard to compute explicit expressions for these coefficients, but I don't see an easy arguments why they can't be negative.</p> <pre><code># Formally compute W var('X N') P = (1+X^(4*N-1))*(1+X^(2*N))*(1-X^(2*N+1)) Q = (1+X^(2*N+1))*(1-X^(2*N+2)) V = P*Q^2*P.diff(X,2)+(P*Q.diff(X))^2-P^2*Q*Q.diff(X,2)-(P.diff(X)*Q)^2 a = 1024*N^2 - 4096/3*N + 320 b = 12288*N^3 - 14336*N^2 + 4864/3*N + 256 W = X^4*(V - (2*N + 1)^2 *(N + 1)^2*X^(24*N-1)*(1-X)^4*(a*X + b*(1-X))) # Check that (1-X)^6 divides W(X) print all(W.diff(X,i)(X=1).polynomial(QQ) == 0 for i in [0..5]) # Compute the coefficients a_ij for the exponents 2ni+j of W. Somewhat # clumsy, as I don't know how to deal with polynomials where exponents # are symbolic expressions. # # l is the list of summands of W l = [z.canonicalize_radical() for z in W.expand().operands()] K.&lt;n&gt; = QQ[] aij = {(i,j):K(0) for i in [0..12] for j in [0..10]} for term in l: c = term(X=1) # get coefficient of term e = (term.diff(X)/c)(X=1) # get exponent of term c = K(c.polynomial(QQ)) # convert c to proper polynomial in n i, j = ZZ(e.diff(N))//2, ZZ(e(N=0)) # Clumsy method to compute pairs (i,j) aij[i,j] += c # Check if coefficients aij[i,j] were correctly computed Wnew = sum(c(n=N)*X^(2*N*i+j) for (i,j),c in aij.items()) print (W-Wnew).canonicalize_radical().is_trivial_zero() def bino(k): # binomial coefficient binom(-6,k)=binom(k+5,5) return prod(k+5-z for z in range(5))/120 # compute the coefficients of W(X)/(1-X)^6 K = K.extend_variables(('y', 'o')) K.inject_variables() for r in [0..12]: for s in [0..10]: d = 1 if s == 10 else 0 # compute coefficient of X^(2nr+s+d*y) f = sum(aij[i,j]*bino(2*n*(r-i)+s+d*y-j) for i in [0..r] for j in [0..s if i == r else 10]) # check non-negativity of the polynomial f if d == 0: # Checks if the coefficient of X^(2nr+s), which is a # polynomial in n, has nonnegative coefficients upon replacing # n with n+12 f = f(n=n+12) if min(f.coefficients()+[0]) &lt; 0: print "False" else: # The coefficient of X^(2nr+10+y), which is a polynomial in n # and y, has to be nonnegative for 0 &lt;= 10+y &lt;= 2n-1, so 10+y # = 2n-1-o for non-negative o. So upon replacing n with # (11+y+o)/2, we get a polynomial in y and o which has to be # nonnegative for all nonnegative y and o. In all but one # case, this holds because the coeffcients are non-negative. f = f(n=(11+y+o)/2) if min(f.coefficients()+[0]) &lt; 0: c = o^2 + 23*o*y + 1360*y^2 + 99*o + 340*y + 1675 print min(c.coefficients()+(c*f).coefficients()) &gt;= 0 </code></pre> Answer 16: <p>This proof is mainly based on some of the suggestions in the early version of this answer. Also, we are building here on the comment by Peter Mueller concerning the polynomial $a$ defined below. It turns out that similar techniques can be used for all other nontrivial polynomials arising in the proof, which allowed us to replace multiple uses of Mathematica command Reduce[] in the last version of this answer by showing that the coefficients of the mentioned relevant polynomials are all nonnegative. I have still retained some use of Reduce[] -- but only to quickly make the routine check for $n\le11$. </p> <p>Note that \begin{equation} G:= F_n''(x) x^{2-2 n} \left(x^{2 n}+1\right)^2 \left(x^{4 n}+x\right)^2 \left(x^{2 n+1}-1\right)^2 \left(x^{2 n+1}+1\right)^2 \left(x^{2 n+2}-1\right)^2 \end{equation} is a polynomial in $n$, $x$, and $y:=x^{2n}$, of degree $2$ in $n$: \begin{equation} G=a+bn+cn^2, \end{equation} where $a,b,c$ are certain polynomials in $x,y$. </p> <p>The check of $G\ge0$ for $n=2,\dots,11$ is straightforward. So, assume $n\ge12$, whence $0&lt;y&lt;x^{24}&lt;x^{12}$. We also always assume that $0&lt;x&lt;1$, $0&lt;y&lt;x^{24}$ , \begin{equation} n=\frac{\ln y}{2\ln x}; \tag{1} \end{equation} the latter relation is of course just another form of the equality $y=x^{2n}$. </p> <p>Note that $0&lt;y&lt;1\ \&amp;\ 0&lt;x&lt;1$ implies $x=1/(1+u)\ \&amp;\ y=1/(1+v)$ for some $u,v\ge0$. As suggested by Peter Mueller, it turns out that \begin{equation} a\; (1 + u)^{10} (1 + v)^{11}|_{x=1/(1 + u), y=1/(1 + v)} \end{equation} is a polynomial in $u,v$ with all coefficients nonnegative. Details of all calculations can be seen in the <a href="https://www.dropbox.com/s/g87a59uzdsb75m5/SolutionM1.nb?dl=0" rel="nofollow noreferrer">Mathematica notebook</a> and its <a href="https://www.dropbox.com/s/k4vlx4brvok099k/SolutionM1.pdf?dl=0" rel="nofollow noreferrer">pdf image</a></p> <p>So, \begin{equation} a\ge0 \end{equation} for all $x,y$ in $(0,1)$. </p> <p>Next, introduce \begin{equation} x_* := \frac{869}{1000},\quad x_0:=\frac{7655}{10000},\quad y_0:= x_0^{24}\approx0.00164,\quad x_1 := \frac{985}{1000},\quad y_1:= x_1^{24}\approx0.696. \end{equation} </p> <p>The following cases/subcases/subsubcases are exhaustive: </p> <p><strong>Case 1: $y\le y_1$.</strong> Then the condition $0&lt;y\le y_1\ \&amp;\ 0&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;x^{24}$ implies $x=x_1/(1+u)\ \&amp;\ y=x^{24}/(1+v)$ or $x=x_1+(1-x_1)/(1+u)\ \&amp;\ y=y_1/(1+v)$ for some $u\ge0$ and $v\ge0$, depending on whether $x&lt;x_1$ or $x\ge x_1$. It turns out that \begin{equation} c\; (1 + u)^{248} (1 + v)^{10}|_{x=x_1/(1+u)\ \&amp;\ y=x^{24}/(1+v)} \end{equation} and \begin{equation} c\; (1 + u)^{10} (1 + v)^{10}|_{x=x_1+(1-x_1)/(1+u)\ \&amp;\ y=y_1/(1+v)} \end{equation} are each a polynomial in $u,v$ with all coefficients nonnegative. So, \begin{equation} c\ge0 \quad\text{in Case 1}. \end{equation} So, for the derivative $G'_n$ of $G$ in $n$, we have \begin{equation}\tag{2} G'_n=b+2cn\ge b+24c=(G'_n)|_{n=12}. \end{equation}</p> <p>Now we have to distinguish two subcases of Case 1, with further subsubcases of Subcase 1.1: </p> <p><strong>Subcase 1.1: $y\le y_1$ and ($y\ge y_0$ or $x\le x_*$).</strong> </p> <p><strong>Subsubcase 1.1.1: $y\le y_1$ and $y\ge y_0$ and $x\ge x_1$.</strong> Then $y\le y_1\ \&amp;\ y\ge y_0\ \&amp;\ x\ge x_1 \ \&amp;\ 0&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;x^{24}$ implies $$x=x_{111}:=x_1+(1-x_1)/(1+u)\ \&amp;\ y=y_{111}:=y_0+(y_1-y_0)/(1+v)$$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=12}\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}} \end{equation} and \begin{equation} G|_{n=12}\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p> <p><strong>Subsubcase 1.1.2: $y\le y_1$ and $y\ge y_0$ and $x&lt;x_1$.</strong> Then $x=x_{112}:=x_0+(x_1-x_0)/(1+u)\ \&amp;\ y=y_{112}:=y_0+(x^{24}-y_0)/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112}, y=y_{112}} \end{equation} and \begin{equation} G|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112},y=y_{112}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p> <p><strong>Subsubcase 1.1.3: $y\le y_1$ and $x\le x_*$.</strong> Then $x=x_{113}:=x_*/(1+u)\ \&amp;\ y=y_{113}:=x^{24}/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}} \end{equation} and \begin{equation} G|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p> <p>So, $G|_{n=12}\ge0$ and $(G'_n)|_{n=12}\ge0$ in all the three subsubcases of Subcase 1.1. In view of $(2)$, we have $G\ge0$ in Subcase 1.1. </p> <p><strong>Subcase 1.2: $y\le y_1$ and $y&lt;y_0$ and $x&gt;x_*$.</strong> Then, of course, the condition $y\le y_1$ is redundant. Let here \begin{equation} \rho(x):=\frac7{2(1-x)}. \end{equation} The condition $x_*&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;y_0$ implies $x=x_{12}:=x_*+(1-x_*)/(1+u)$ and $y=y_{12}:=y_0/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=\rho(x)/2}\; u\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}} \end{equation} and \begin{equation} G|_{n=\rho(x)/2}\; u^2\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. So, similarly to Subcase 1.1, for all $n\ge\rho(x)/2$ we get $G\ge G|_{n=\rho(x)/2}\ge0$. In Subcase 1.2, it remains to note that for our particular $n$, as in $(1)$, we indeed have $n\ge\rho(x)/2$. This follows because for $y&lt;y_0$ and $x&gt;x_*$ \begin{equation} \ln\frac1y&gt;\ln\frac1{y_0}&gt;1.08\frac7{2(1-x_*)}\,\ln\frac1{x_*} &gt;\frac7{2(1-x)}\,\ln\frac1{x}=\rho(x)\ln\frac1{x}. \end{equation} So, $G\ge0$ in Subcase 1.2 as well. </p> <p>It remains to consider </p> <p><strong>Case 2: $y&gt;y_1$.</strong> Then condition $y&gt;y_1\ \&amp;\ 0&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;x^{24}[&lt;x^{12}]$ implies $x=x_2:=x_1+(1-x_1)/(1+u),y=y_2:=y_1+(x^{12}-y_1)/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} b\;(1 + u)^{140} (1 + v)^{11}|_{x=x_{2},y=y_{2}} \end{equation} is also a polynomial in $u,v$ with all coefficients nonnegative. Hence, $b\ge0$ in Case 2. So, if $c\ge0$, then trivially $G\ge0$ (since $a\ge0$ always). So, without loss of generality, $c&lt;0$ and hence $G$ is concave in $n$. So, it is enough to bracket the $n$ as in $(1)$ between some simple rational expressions $n_1$ and $n_2$ such that $G|_{n=n_1}\ge0$ and $G|_{n=n_2}\ge0$. </p> <p>In Case 2, $y$ is "close" to $1$ and hence, in view of inequality $y&lt;x^{24}$, so is $x$: $x&gt;y_1^{1/12}=x_1=0.985$. So, it is a bit more convenient here to change variables $x, y$ to "small variables" $u:=1-x$ and $v:=1-y$. Then $0&lt;v&lt;1-y_1\approx0.304$ and $1-u=x&gt;y^{1/24}=(1-v)^{1/24}&gt;1-v/3$, whence $u&lt;v/3$. Let now $t:=u/v$, so that $0&lt;t&lt;1/3$. It follows, in Case 2, that $t=(1/3)/(1+r)$ and $v=(1-y_1)/(1+s)$ for some $r\ge0$ and $s\ge0$. </p> <p>The mentioned brackets $n_1$ and $n_2$ are respectively $\ell_1/2$ and $\ell_2/2$, where<br> \begin{equation} \ell_1:=\frac1t,\quad\ell_2:=\frac{(2 - v) (1 - t v)}{t (1 - v) (2 - t v)}. \end{equation} It is not hard to see (details on this are in the same Mathematica notebook) that then indeed $n_1 &lt; n &lt; n_2$. It remains to show that $G|_{n=n_1}\ge0$ and $G|_{n=n_2}\ge0$. </p> <p>It turns out that \begin{equation} (G'_n)|_{n=\ell_1/2;\,x=1-tv,\,y=1-v;\,t=(1/3)/(1+r),\,v=(1-y_1)/(1+s)}\; (1 + r)^{10} (1 + s)^{19} \end{equation} and \begin{equation} (G'_n)|_{n=\ell_2/2;\,x=1-tv,\,y=1-v;\,t=(1/3)/(1+r),\,v=(1-y_1)/(1+s)}\; (1 + r)^{10} (1 + s)^{19}\; q \end{equation} are each a polynomial in $r,s$ with all coefficients nonnegative, where \begin{multline*} q:= (1 + r)^{10} (1 + s)^{19} \\ \times(11673186598630578538556565100133681446610566511878526881 \\ + 16777216000000000000000000000000000000000000000000000000 s)^2 \\ \times\big(31853088866210192846185521700044560482203522170626175627 \\ + 33554432000000000000000000000000000000000000000000000000 (r + s + r s)\big)^2. \end{multline*}</p> <p>The proof is complete. </p> Answer 17: <p>I sketch a method that should show $F_n$ is convex for $n$ sufficiently large.</p> <p>A Taylor expansion gives that $F_n(x) = x^{2n} (1-x)^2 + O(x^{4n-1})$, and so $$F_n''(x) = 2x^{2n-2}[2n^2(1-x)^2 + x^2 - n(1-x)(1+3x)] + O(x^{4n-3}).$$ This should show that $F_n''(x) &gt; 0$ for $n$ large, provided $x &lt; 1-\frac{C}{n}$ for some fixed $C&gt;0$.</p> <p>On the other hand, according to Mathematica, $F_n''(1) = 4n^2 - \frac{16}{3} n + \frac{5}{4}$, which has its largest root at $\frac{8 + \sqrt{19}}{12} \approx 1.02991$. (By the way, there doesn't seem to be any reason that $n$ should be restricted to be an integer, and this is in line with some plots I made indicating $F_n(x)$ is convex for $n &gt; 1.03$.) Moreover, Mathematica gives $$\lim_{n\rightarrow \infty} n^{-2}F_n''(1-\frac{y}{n}) = \frac{16 e^{4y}}{(1+e^{4y})^2} &gt; 0.$$</p> <p>Putting these two pieces together should be able to show $F_n(x)$ is convex on $(0,1)$ for $n$ sufficiently large. Unfortunately, it might be very painful to work out all the error terms explicitly.</p> Answer 18: <p>Not an answer, just a plot of $F_n(x)$ for $n=2,\ldots,10$: <hr /> &nbsp; &nbsp; &nbsp; &nbsp; &nbsp; <a href="https://i.stack.imgur.com/jzWwY.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/jzWwY.jpg" alt="Fn(x)"></a> <hr /> &nbsp; &nbsp; &nbsp; &nbsp; &nbsp; <a href="https://i.stack.imgur.com/c8gFL.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/c8gFL.jpg" alt="Fn(x)2"></a></p> <hr /> Answer 19: <p>This is not an answer, just a reformulation into something which looks perhaps tractable (along the lines of the comment by G. Paseman). Define, for $m\geq 1$ and $x\in (0,1)$, $$g_m(x):=\frac{d^2}{dx^2}[\log(1+x^m)]=\frac{x^{m-2}(m(m-1)-mx^m)}{(1+x^m)^2}$$ and $$h_m(x):=\frac{d^2}{dx^2}[\log(1-x^m)]=-\frac{x^{m-2}(m(m-1)+mx^m)}{(1-x^m)^2}.$$ Then, for $n\geq 2$, $$F_n''(x)=g_{4n-1}(x)+g_{2n}(x)+h_{2n+1}(x)-g_{2n+1}(x)-h_{2n+2}(x).$$ </p> <p>Two observations</p> <ul> <li>The contribution of the first, second, and last terms is non-negative for all $n\geq 2$ and $x\in(0,1)$, while that of the remaining terms is non-positive.</li> <li>It would therefore suffice to show the following: for $n$ and $x$ as above, $$g_{2n+1}(x)-h_{2n+1}(x)\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{2n+2}(x).$$</li> </ul> <p><strong>Added</strong>: Since I apparently cannot add a comment, I would just note that while it is certainly (always) possible I overlook something, I have just double checked the algebra with maple, and believe that the expressions are correct (there doesn't seem to be a missing factor of m). Another note (in addition to the symmetry observed in the comment by Paseman) is the (trivial) identity $h_{2m}=g_m+h_m$, leading to another form for the desired bound: $$g_{2n+1}(x)+g_{n+1}(x)-h_{2n+1}(x)\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{n+1}(x).$$</p> Answer 20: <p>This is only a comment to @DimaPasechnik, but I cannot put the picture in a comment. The surface to the right of the $x=y$ is a plot of Dima's function (barring mistakes); clearly not convex.</p> <p><a href="https://i.stack.imgur.com/7DSRU.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/7DSRU.jpg" alt="enter image description here"></a></p> Answer 21: <p>I conjecture a stronger statement: that the function $F(x,y)$ of two variables $x,y$ obtained from $F_n(x)$ by substituting $y$ for $x^{2n}$ is convex for $0&lt;y&lt;x&lt;0$. </p> <p>EDIT2: this conjecture is wrong as stated. See e.g. the plot in the answer by Yaakov Baruch, or rotate the plot in Sage...</p> <p>While this reduces to checking positive definiteness of a matrix of bivariate polynomials of degree about 20, obtained from the Hessian of $F$ by clearing common (positive) denominator, this should be possible to make to work by standard real algebraic geometry tools. Here is an (ugly) plot of $F$:</p> <p><a href="https://i.stack.imgur.com/l9j8a.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/l9j8a.jpg" alt="plot on [0,1]x[0,1]"></a> obtained by exporting the plot from jmol in Sage(math) </p> <pre><code>sage: var('x y') sage: F(x,y)=log((x+y^2)*(1+y)*(1-x*y)/(x*(1+x*y)*(1-x^2*y))) sage: plot3d(F,[0,1],[0,1]) </code></pre> <p>EDIT: comments say that as $y\to 0+$, the Hessian becomes indefinite. That is, for my original claim to hold, one need to assume $y&gt;\epsilon&gt;0$, with $\epsilon$ possibly depending upon $x$. </p> Answer 22: <p>Fix $n$, and let $V(x)$ be the polynomial from the question, for which we want to show that it is positive on the open interval $(0,1)$. Set \begin{align*} a(n) &amp;= 1024n^2 - 4096/3n + 320\\ b(n) &amp;= 12288n^3 - 14336n^2 + 4864/3n + 256\\ W(x) &amp;= x^4\left(V(x)-(n+1)^2(2n+1)^2x^{24n-1}(1-x)^4(a(n)x + b(n)(1-x))\right). \end{align*}</p> <p>For all $n\ge12$ we claim that $(1-x)^6$ divides $W(x)$, and that all the coefficients of $W(x)/(1-x)^6$ are nonnegative. In particular, $W(x)$ is positive for $x&gt;0$. As $a(n)$ and $b(n)$ is positive for all $n\ge1$, we see that $V(x)$ is positive for all $0&lt;x&lt;1$, provided that $n\ge12$. The cases $2\le n\le 11$ are easily checked directly, for instance by noting that the coefficients of $(1+x)^{\deg V}V(1/(1+x))$ are positive.</p> <p>I'm not sure about the easiest kind to prove the assertion about the coefficients of $W(x)/(1-x)^6$. We compute them via \begin{equation} \frac{W(x)}{(1-x)^6}=W(x)\sum_{k\ge0}\binom{k+5}{5}x^k. \end{equation}</p> <p>All what follows is assisted/confirmed by the Sage code below.</p> <p>The exponents of $W(x)$ have the form $2ni+j$ for $0\le i\le 12$ and $0\le j\le 10$. Let $a_{i,j}$ be the corresponding coefficient. It is a polynomial in $n$.</p> <p>We have \begin{equation} \frac{W(x)}{(1-x)^6} = \sum_{i,j}a_{i,j}(n)x^{2ni+j}\sum_k\binom{k+5}{5}x^k. \end{equation}</p> <p>With $r\ge0$ and $0\le s\le 2n-1$, the coefficient of $x^{2nr+s}$ in $W(x)\frac{1}{(1-x)^6}$ is the sum of all $a_{i,j}(n)\cdot\binom{nr+s-ni-j+y}{5}$ for all pairs $(i,j)$ where $i&lt;r$ and $0\le j\le 10$, or $i=r$ and $j\le s$.</p> <p>We distinguish the cases $0\le s\le 9$ from the cases $s\ge10$. In these latter cases, we write $s=10+y$. Recall that $s\le 2n-1$, so $10+y=2n-1-o$ for a nonnegative $o$. Upon replacing $n$ with $(11+y+1)/2$, we get polynomials in $n$ and $o$.</p> <p>In the former cases, the coefficients are polynomials in $n$, which have positive coefficients upon replacing $n$ with $n+12$. So these coefficients are positive for all $n\ge12$. </p> <p>In the latter cases it turns out that in all but one case the coefficients are positive. So for nonnegative $y$ and $o$, the values are nonnegative.</p> <p>For the single exception we see that if we multiply it with a suitable polynomial, the resulting coefficients are positive.</p> <p><em>Remark (answering Jason's question):</em> It is a known fact that a polynomial $V$ is positive on $(0,1)$ if and only if it is a nonnegative linear combination of polynomials $x^i(1-x)^j$. The problem is that it may involve terms where $i+j$ is bigger than $\deg V$. (This doesn't happen here, though.) I used an LP solver to play a little with $V$ and $\frac{V}{(1-x^2)^4x^{2n-2}}$. From that a pattern showed up which lead to a solution.</p> <p>By the way, $V$ actually seems to be a nonnegative linear combination of $x^i(1-x)^{\deg V-i}$. This is equivalent to say that all the coefficients of $(1+x)^{\deg V}V(1/(1+x))$ are nonnegative. It is not hard to compute explicit expressions for these coefficients, but I don't see an easy arguments why they can't be negative.</p> <pre><code># Formally compute W var('X N') P = (1+X^(4*N-1))*(1+X^(2*N))*(1-X^(2*N+1)) Q = (1+X^(2*N+1))*(1-X^(2*N+2)) V = P*Q^2*P.diff(X,2)+(P*Q.diff(X))^2-P^2*Q*Q.diff(X,2)-(P.diff(X)*Q)^2 a = 1024*N^2 - 4096/3*N + 320 b = 12288*N^3 - 14336*N^2 + 4864/3*N + 256 W = X^4*(V - (2*N + 1)^2 *(N + 1)^2*X^(24*N-1)*(1-X)^4*(a*X + b*(1-X))) # Check that (1-X)^6 divides W(X) print all(W.diff(X,i)(X=1).polynomial(QQ) == 0 for i in [0..5]) # Compute the coefficients a_ij for the exponents 2ni+j of W. Somewhat # clumsy, as I don't know how to deal with polynomials where exponents # are symbolic expressions. # # l is the list of summands of W l = [z.canonicalize_radical() for z in W.expand().operands()] K.&lt;n&gt; = QQ[] aij = {(i,j):K(0) for i in [0..12] for j in [0..10]} for term in l: c = term(X=1) # get coefficient of term e = (term.diff(X)/c)(X=1) # get exponent of term c = K(c.polynomial(QQ)) # convert c to proper polynomial in n i, j = ZZ(e.diff(N))//2, ZZ(e(N=0)) # Clumsy method to compute pairs (i,j) aij[i,j] += c # Check if coefficients aij[i,j] were correctly computed Wnew = sum(c(n=N)*X^(2*N*i+j) for (i,j),c in aij.items()) print (W-Wnew).canonicalize_radical().is_trivial_zero() def bino(k): # binomial coefficient binom(-6,k)=binom(k+5,5) return prod(k+5-z for z in range(5))/120 # compute the coefficients of W(X)/(1-X)^6 K = K.extend_variables(('y', 'o')) K.inject_variables() for r in [0..12]: for s in [0..10]: d = 1 if s == 10 else 0 # compute coefficient of X^(2nr+s+d*y) f = sum(aij[i,j]*bino(2*n*(r-i)+s+d*y-j) for i in [0..r] for j in [0..s if i == r else 10]) # check non-negativity of the polynomial f if d == 0: # Checks if the coefficient of X^(2nr+s), which is a # polynomial in n, has nonnegative coefficients upon replacing # n with n+12 f = f(n=n+12) if min(f.coefficients()+[0]) &lt; 0: print "False" else: # The coefficient of X^(2nr+10+y), which is a polynomial in n # and y, has to be nonnegative for 0 &lt;= 10+y &lt;= 2n-1, so 10+y # = 2n-1-o for non-negative o. So upon replacing n with # (11+y+o)/2, we get a polynomial in y and o which has to be # nonnegative for all nonnegative y and o. In all but one # case, this holds because the coeffcients are non-negative. f = f(n=(11+y+o)/2) if min(f.coefficients()+[0]) &lt; 0: c = o^2 + 23*o*y + 1360*y^2 + 99*o + 340*y + 1675 print min(c.coefficients()+(c*f).coefficients()) &gt;= 0 </code></pre> Answer 23: <p>This proof is mainly based on some of the suggestions in the early version of this answer. Also, we are building here on the comment by Peter Mueller concerning the polynomial $a$ defined below. It turns out that similar techniques can be used for all other nontrivial polynomials arising in the proof, which allowed us to replace multiple uses of Mathematica command Reduce[] in the last version of this answer by showing that the coefficients of the mentioned relevant polynomials are all nonnegative. I have still retained some use of Reduce[] -- but only to quickly make the routine check for $n\le11$. </p> <p>Note that \begin{equation} G:= F_n''(x) x^{2-2 n} \left(x^{2 n}+1\right)^2 \left(x^{4 n}+x\right)^2 \left(x^{2 n+1}-1\right)^2 \left(x^{2 n+1}+1\right)^2 \left(x^{2 n+2}-1\right)^2 \end{equation} is a polynomial in $n$, $x$, and $y:=x^{2n}$, of degree $2$ in $n$: \begin{equation} G=a+bn+cn^2, \end{equation} where $a,b,c$ are certain polynomials in $x,y$. </p> <p>The check of $G\ge0$ for $n=2,\dots,11$ is straightforward. So, assume $n\ge12$, whence $0&lt;y&lt;x^{24}&lt;x^{12}$. We also always assume that $0&lt;x&lt;1$, $0&lt;y&lt;x^{24}$ , \begin{equation} n=\frac{\ln y}{2\ln x}; \tag{1} \end{equation} the latter relation is of course just another form of the equality $y=x^{2n}$. </p> <p>Note that $0&lt;y&lt;1\ \&amp;\ 0&lt;x&lt;1$ implies $x=1/(1+u)\ \&amp;\ y=1/(1+v)$ for some $u,v\ge0$. As suggested by Peter Mueller, it turns out that \begin{equation} a\; (1 + u)^{10} (1 + v)^{11}|_{x=1/(1 + u), y=1/(1 + v)} \end{equation} is a polynomial in $u,v$ with all coefficients nonnegative. Details of all calculations can be seen in the <a href="https://www.dropbox.com/s/g87a59uzdsb75m5/SolutionM1.nb?dl=0" rel="nofollow noreferrer">Mathematica notebook</a> and its <a href="https://www.dropbox.com/s/k4vlx4brvok099k/SolutionM1.pdf?dl=0" rel="nofollow noreferrer">pdf image</a></p> <p>So, \begin{equation} a\ge0 \end{equation} for all $x,y$ in $(0,1)$. </p> <p>Next, introduce \begin{equation} x_* := \frac{869}{1000},\quad x_0:=\frac{7655}{10000},\quad y_0:= x_0^{24}\approx0.00164,\quad x_1 := \frac{985}{1000},\quad y_1:= x_1^{24}\approx0.696. \end{equation} </p> <p>The following cases/subcases/subsubcases are exhaustive: </p> <p><strong>Case 1: $y\le y_1$.</strong> Then the condition $0&lt;y\le y_1\ \&amp;\ 0&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;x^{24}$ implies $x=x_1/(1+u)\ \&amp;\ y=x^{24}/(1+v)$ or $x=x_1+(1-x_1)/(1+u)\ \&amp;\ y=y_1/(1+v)$ for some $u\ge0$ and $v\ge0$, depending on whether $x&lt;x_1$ or $x\ge x_1$. It turns out that \begin{equation} c\; (1 + u)^{248} (1 + v)^{10}|_{x=x_1/(1+u)\ \&amp;\ y=x^{24}/(1+v)} \end{equation} and \begin{equation} c\; (1 + u)^{10} (1 + v)^{10}|_{x=x_1+(1-x_1)/(1+u)\ \&amp;\ y=y_1/(1+v)} \end{equation} are each a polynomial in $u,v$ with all coefficients nonnegative. So, \begin{equation} c\ge0 \quad\text{in Case 1}. \end{equation} So, for the derivative $G'_n$ of $G$ in $n$, we have \begin{equation}\tag{2} G'_n=b+2cn\ge b+24c=(G'_n)|_{n=12}. \end{equation}</p> <p>Now we have to distinguish two subcases of Case 1, with further subsubcases of Subcase 1.1: </p> <p><strong>Subcase 1.1: $y\le y_1$ and ($y\ge y_0$ or $x\le x_*$).</strong> </p> <p><strong>Subsubcase 1.1.1: $y\le y_1$ and $y\ge y_0$ and $x\ge x_1$.</strong> Then $y\le y_1\ \&amp;\ y\ge y_0\ \&amp;\ x\ge x_1 \ \&amp;\ 0&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;x^{24}$ implies $$x=x_{111}:=x_1+(1-x_1)/(1+u)\ \&amp;\ y=y_{111}:=y_0+(y_1-y_0)/(1+v)$$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=12}\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}} \end{equation} and \begin{equation} G|_{n=12}\; (1 + u)^{10} (1 + v)^{11}|_{x=x_{111}, y=y_{111}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p> <p><strong>Subsubcase 1.1.2: $y\le y_1$ and $y\ge y_0$ and $x&lt;x_1$.</strong> Then $x=x_{112}:=x_0+(x_1-x_0)/(1+u)\ \&amp;\ y=y_{112}:=y_0+(x^{24}-y_0)/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112}, y=y_{112}} \end{equation} and \begin{equation} G|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{112},y=y_{112}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p> <p><strong>Subsubcase 1.1.3: $y\le y_1$ and $x\le x_*$.</strong> Then $x=x_{113}:=x_*/(1+u)\ \&amp;\ y=y_{113}:=x^{24}/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}} \end{equation} and \begin{equation} G|_{n=12}\; (1 + u)^{272} (1 + v)^{11}|_{x=x_{113},y=y_{113}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. </p> <p>So, $G|_{n=12}\ge0$ and $(G'_n)|_{n=12}\ge0$ in all the three subsubcases of Subcase 1.1. In view of $(2)$, we have $G\ge0$ in Subcase 1.1. </p> <p><strong>Subcase 1.2: $y\le y_1$ and $y&lt;y_0$ and $x&gt;x_*$.</strong> Then, of course, the condition $y\le y_1$ is redundant. Let here \begin{equation} \rho(x):=\frac7{2(1-x)}. \end{equation} The condition $x_*&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;y_0$ implies $x=x_{12}:=x_*+(1-x_*)/(1+u)$ and $y=y_{12}:=y_0/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} (G'_n)|_{n=\rho(x)/2}\; u\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}} \end{equation} and \begin{equation} G|_{n=\rho(x)/2}\; u^2\,(1 + u)^{10} (1 + v)^{11}|_{x=x_{12},y=y_{12}} \end{equation} are each, again, a polynomial in $u,v$ with all coefficients nonnegative. So, similarly to Subcase 1.1, for all $n\ge\rho(x)/2$ we get $G\ge G|_{n=\rho(x)/2}\ge0$. In Subcase 1.2, it remains to note that for our particular $n$, as in $(1)$, we indeed have $n\ge\rho(x)/2$. This follows because for $y&lt;y_0$ and $x&gt;x_*$ \begin{equation} \ln\frac1y&gt;\ln\frac1{y_0}&gt;1.08\frac7{2(1-x_*)}\,\ln\frac1{x_*} &gt;\frac7{2(1-x)}\,\ln\frac1{x}=\rho(x)\ln\frac1{x}. \end{equation} So, $G\ge0$ in Subcase 1.2 as well. </p> <p>It remains to consider </p> <p><strong>Case 2: $y&gt;y_1$.</strong> Then condition $y&gt;y_1\ \&amp;\ 0&lt;x&lt;1\ \&amp;\ 0&lt;y&lt;x^{24}[&lt;x^{12}]$ implies $x=x_2:=x_1+(1-x_1)/(1+u),y=y_2:=y_1+(x^{12}-y_1)/(1+v)$ for some $u\ge0$ and $v\ge0$. It turns out that \begin{equation} b\;(1 + u)^{140} (1 + v)^{11}|_{x=x_{2},y=y_{2}} \end{equation} is also a polynomial in $u,v$ with all coefficients nonnegative. Hence, $b\ge0$ in Case 2. So, if $c\ge0$, then trivially $G\ge0$ (since $a\ge0$ always). So, without loss of generality, $c&lt;0$ and hence $G$ is concave in $n$. So, it is enough to bracket the $n$ as in $(1)$ between some simple rational expressions $n_1$ and $n_2$ such that $G|_{n=n_1}\ge0$ and $G|_{n=n_2}\ge0$. </p> <p>In Case 2, $y$ is "close" to $1$ and hence, in view of inequality $y&lt;x^{24}$, so is $x$: $x&gt;y_1^{1/12}=x_1=0.985$. So, it is a bit more convenient here to change variables $x, y$ to "small variables" $u:=1-x$ and $v:=1-y$. Then $0&lt;v&lt;1-y_1\approx0.304$ and $1-u=x&gt;y^{1/24}=(1-v)^{1/24}&gt;1-v/3$, whence $u&lt;v/3$. Let now $t:=u/v$, so that $0&lt;t&lt;1/3$. It follows, in Case 2, that $t=(1/3)/(1+r)$ and $v=(1-y_1)/(1+s)$ for some $r\ge0$ and $s\ge0$. </p> <p>The mentioned brackets $n_1$ and $n_2$ are respectively $\ell_1/2$ and $\ell_2/2$, where<br> \begin{equation} \ell_1:=\frac1t,\quad\ell_2:=\frac{(2 - v) (1 - t v)}{t (1 - v) (2 - t v)}. \end{equation} It is not hard to see (details on this are in the same Mathematica notebook) that then indeed $n_1 &lt; n &lt; n_2$. It remains to show that $G|_{n=n_1}\ge0$ and $G|_{n=n_2}\ge0$. </p> <p>It turns out that \begin{equation} (G'_n)|_{n=\ell_1/2;\,x=1-tv,\,y=1-v;\,t=(1/3)/(1+r),\,v=(1-y_1)/(1+s)}\; (1 + r)^{10} (1 + s)^{19} \end{equation} and \begin{equation} (G'_n)|_{n=\ell_2/2;\,x=1-tv,\,y=1-v;\,t=(1/3)/(1+r),\,v=(1-y_1)/(1+s)}\; (1 + r)^{10} (1 + s)^{19}\; q \end{equation} are each a polynomial in $r,s$ with all coefficients nonnegative, where \begin{multline*} q:= (1 + r)^{10} (1 + s)^{19} \\ \times(11673186598630578538556565100133681446610566511878526881 \\ + 16777216000000000000000000000000000000000000000000000000 s)^2 \\ \times\big(31853088866210192846185521700044560482203522170626175627 \\ + 33554432000000000000000000000000000000000000000000000000 (r + s + r s)\big)^2. \end{multline*}</p> <p>The proof is complete. </p> Answer 24: <p>I sketch a method that should show $F_n$ is convex for $n$ sufficiently large.</p> <p>A Taylor expansion gives that $F_n(x) = x^{2n} (1-x)^2 + O(x^{4n-1})$, and so $$F_n''(x) = 2x^{2n-2}[2n^2(1-x)^2 + x^2 - n(1-x)(1+3x)] + O(x^{4n-3}).$$ This should show that $F_n''(x) &gt; 0$ for $n$ large, provided $x &lt; 1-\frac{C}{n}$ for some fixed $C&gt;0$.</p> <p>On the other hand, according to Mathematica, $F_n''(1) = 4n^2 - \frac{16}{3} n + \frac{5}{4}$, which has its largest root at $\frac{8 + \sqrt{19}}{12} \approx 1.02991$. (By the way, there doesn't seem to be any reason that $n$ should be restricted to be an integer, and this is in line with some plots I made indicating $F_n(x)$ is convex for $n &gt; 1.03$.) Moreover, Mathematica gives $$\lim_{n\rightarrow \infty} n^{-2}F_n''(1-\frac{y}{n}) = \frac{16 e^{4y}}{(1+e^{4y})^2} &gt; 0.$$</p> <p>Putting these two pieces together should be able to show $F_n(x)$ is convex on $(0,1)$ for $n$ sufficiently large. Unfortunately, it might be very painful to work out all the error terms explicitly.</p> Answer 25: <p>Not an answer, just a plot of $F_n(x)$ for $n=2,\ldots,10$: <hr /> &nbsp; &nbsp; &nbsp; &nbsp; &nbsp; <a href="https://i.stack.imgur.com/jzWwY.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/jzWwY.jpg" alt="Fn(x)"></a> <hr /> &nbsp; &nbsp; &nbsp; &nbsp; &nbsp; <a href="https://i.stack.imgur.com/c8gFL.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/c8gFL.jpg" alt="Fn(x)2"></a></p> <hr /> Answer 26: <p>This is not an answer, just a reformulation into something which looks perhaps tractable (along the lines of the comment by G. Paseman). Define, for $m\geq 1$ and $x\in (0,1)$, $$g_m(x):=\frac{d^2}{dx^2}[\log(1+x^m)]=\frac{x^{m-2}(m(m-1)-mx^m)}{(1+x^m)^2}$$ and $$h_m(x):=\frac{d^2}{dx^2}[\log(1-x^m)]=-\frac{x^{m-2}(m(m-1)+mx^m)}{(1-x^m)^2}.$$ Then, for $n\geq 2$, $$F_n''(x)=g_{4n-1}(x)+g_{2n}(x)+h_{2n+1}(x)-g_{2n+1}(x)-h_{2n+2}(x).$$ </p> <p>Two observations</p> <ul> <li>The contribution of the first, second, and last terms is non-negative for all $n\geq 2$ and $x\in(0,1)$, while that of the remaining terms is non-positive.</li> <li>It would therefore suffice to show the following: for $n$ and $x$ as above, $$g_{2n+1}(x)-h_{2n+1}(x)\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{2n+2}(x).$$</li> </ul> <p><strong>Added</strong>: Since I apparently cannot add a comment, I would just note that while it is certainly (always) possible I overlook something, I have just double checked the algebra with maple, and believe that the expressions are correct (there doesn't seem to be a missing factor of m). Another note (in addition to the symmetry observed in the comment by Paseman) is the (trivial) identity $h_{2m}=g_m+h_m$, leading to another form for the desired bound: $$g_{2n+1}(x)+g_{n+1}(x)-h_{2n+1}(x)\leq_? g_{4n-1}(x)+g_{2n}(x)-h_{n+1}(x).$$</p> Answer 27: <p>This is only a comment to @DimaPasechnik, but I cannot put the picture in a comment. The surface to the right of the $x=y$ is a plot of Dima's function (barring mistakes); clearly not convex.</p> <p><a href="https://i.stack.imgur.com/7DSRU.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/7DSRU.jpg" alt="enter image description here"></a></p> Answer 28: <p>I conjecture a stronger statement: that the function $F(x,y)$ of two variables $x,y$ obtained from $F_n(x)$ by substituting $y$ for $x^{2n}$ is convex for $0&lt;y&lt;x&lt;0$. </p> <p>EDIT2: this conjecture is wrong as stated. See e.g. the plot in the answer by Yaakov Baruch, or rotate the plot in Sage...</p> <p>While this reduces to checking positive definiteness of a matrix of bivariate polynomials of degree about 20, obtained from the Hessian of $F$ by clearing common (positive) denominator, this should be possible to make to work by standard real algebraic geometry tools. Here is an (ugly) plot of $F$:</p> <p><a href="https://i.stack.imgur.com/l9j8a.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/l9j8a.jpg" alt="plot on [0,1]x[0,1]"></a> obtained by exporting the plot from jmol in Sage(math) </p> <pre><code>sage: var('x y') sage: F(x,y)=log((x+y^2)*(1+y)*(1-x*y)/(x*(1+x*y)*(1-x^2*y))) sage: plot3d(F,[0,1],[0,1]) </code></pre> <p>EDIT: comments say that as $y\to 0+$, the Hessian becomes indefinite. That is, for my original claim to hold, one need to assume $y&gt;\epsilon&gt;0$, with $\epsilon$ possibly depending upon $x$. </p>
Could the Riemann zeta function be a solution for a known differential equation?
260,196
43
https://mathoverflow.net/questions/260196/could-the-riemann-zeta-function-be-a-solution-for-a-known-differential-equation
<p><a href="https://en.wikipedia.org/wiki/Riemann_zeta_function">Riemann zeta function</a> is a function of complex variable $s$ that analytically continous the sum of Dirichlet series .defined as :$$\zeta(s)=\sum_{n=1}^{\infty}\displaystyle \frac{1}{n^s} $$ for when the real part is greater than $1$.</p> <blockquote> <blockquote> <p><strong>My question here is:</strong> Could the Riemann zeta function be a solution for a known differential equation?</p> </blockquote> </blockquote> <p><strong>Note:</strong> I would like if there is a paper or ref show that zeta function presented a solution for known Differential equation.</p>
[ { "answer_id": 260198, "score": 45, "text": "<p>When posed properly, a long-standing open problem, but in the form you ask:</p>\n\n<p><em>Robert A. Van Gorder</em>, MR 3276353 <a href=\"http://dx.doi.org/10.1016/j.jnt.2014.08.013\"><strong>Does the Riemann zeta function satisfy a differential equation?</strong></a>, <em>J. Number Theory</em> <strong>147</strong> (2015), 778--788.</p>\n" }, { "answer_id": 260205, "score": 21, "text": "<p>The fact that $\\zeta$ satisfy no algebraic differential equation is due to its famous relation with the Gamma function which was proved by Hölder not to satisfy such an equation.</p>\n\n<p>Detailed answer can be found <a href=\"https://math.stackexchange.com/questions/1423720/is-there-any-proof-that-the-riemann-zeta-function-is-not-elementary\">here</a> with five (commented and linked) references. </p>\n\n<p>On the other hand, this function is linked with many other transcendental special functions like polylogarithms which satisfy <a href=\"https://arxiv.org/abs/1101.4497v1\" rel=\"nofollow noreferrer\">Fuschian non commutative differential equations</a>. </p>\n" }, { "answer_id": 260208, "score": 21, "text": "<p>The Riemann zeta function is \"hypertranscendental\" in the sense shown <a href=\"https://en.wikipedia.org/wiki/Hypertranscendental_function\">HERE</a></p>\n\n<p>It is not the solution $y(x)$ of a differential equation of the form\n$$\nF(x,y,y',y'',\\dots,y^{(n)})=0\n$$\nwhere $F$ is a polynomial (with constant coefficients).</p>\n" }, { "answer_id": 260198, "score": 45, "text": "<p>When posed properly, a long-standing open problem, but in the form you ask:</p>\n\n<p><em>Robert A. Van Gorder</em>, MR 3276353 <a href=\"http://dx.doi.org/10.1016/j.jnt.2014.08.013\"><strong>Does the Riemann zeta function satisfy a differential equation?</strong></a>, <em>J. Number Theory</em> <strong>147</strong> (2015), 778--788.</p>\n" }, { "answer_id": 260205, "score": 21, "text": "<p>The fact that $\\zeta$ satisfy no algebraic differential equation is due to its famous relation with the Gamma function which was proved by Hölder not to satisfy such an equation.</p>\n\n<p>Detailed answer can be found <a href=\"https://math.stackexchange.com/questions/1423720/is-there-any-proof-that-the-riemann-zeta-function-is-not-elementary\">here</a> with five (commented and linked) references. </p>\n\n<p>On the other hand, this function is linked with many other transcendental special functions like polylogarithms which satisfy <a href=\"https://arxiv.org/abs/1101.4497v1\" rel=\"nofollow noreferrer\">Fuschian non commutative differential equations</a>. </p>\n" }, { "answer_id": 260208, "score": 21, "text": "<p>The Riemann zeta function is \"hypertranscendental\" in the sense shown <a href=\"https://en.wikipedia.org/wiki/Hypertranscendental_function\">HERE</a></p>\n\n<p>It is not the solution $y(x)$ of a differential equation of the form\n$$\nF(x,y,y',y'',\\dots,y^{(n)})=0\n$$\nwhere $F$ is a polynomial (with constant coefficients).</p>\n" }, { "answer_id": 260198, "score": 45, "text": "<p>When posed properly, a long-standing open problem, but in the form you ask:</p>\n\n<p><em>Robert A. Van Gorder</em>, MR 3276353 <a href=\"http://dx.doi.org/10.1016/j.jnt.2014.08.013\"><strong>Does the Riemann zeta function satisfy a differential equation?</strong></a>, <em>J. Number Theory</em> <strong>147</strong> (2015), 778--788.</p>\n" }, { "answer_id": 260205, "score": 21, "text": "<p>The fact that $\\zeta$ satisfy no algebraic differential equation is due to its famous relation with the Gamma function which was proved by Hölder not to satisfy such an equation.</p>\n\n<p>Detailed answer can be found <a href=\"https://math.stackexchange.com/questions/1423720/is-there-any-proof-that-the-riemann-zeta-function-is-not-elementary\">here</a> with five (commented and linked) references. </p>\n\n<p>On the other hand, this function is linked with many other transcendental special functions like polylogarithms which satisfy <a href=\"https://arxiv.org/abs/1101.4497v1\" rel=\"nofollow noreferrer\">Fuschian non commutative differential equations</a>. </p>\n" }, { "answer_id": 260208, "score": 21, "text": "<p>The Riemann zeta function is \"hypertranscendental\" in the sense shown <a href=\"https://en.wikipedia.org/wiki/Hypertranscendental_function\">HERE</a></p>\n\n<p>It is not the solution $y(x)$ of a differential equation of the form\n$$\nF(x,y,y',y'',\\dots,y^{(n)})=0\n$$\nwhere $F$ is a polynomial (with constant coefficients).</p>\n" }, { "answer_id": 260198, "score": 45, "text": "<p>When posed properly, a long-standing open problem, but in the form you ask:</p>\n\n<p><em>Robert A. Van Gorder</em>, MR 3276353 <a href=\"http://dx.doi.org/10.1016/j.jnt.2014.08.013\"><strong>Does the Riemann zeta function satisfy a differential equation?</strong></a>, <em>J. Number Theory</em> <strong>147</strong> (2015), 778--788.</p>\n" }, { "answer_id": 260205, "score": 21, "text": "<p>The fact that $\\zeta$ satisfy no algebraic differential equation is due to its famous relation with the Gamma function which was proved by Hölder not to satisfy such an equation.</p>\n\n<p>Detailed answer can be found <a href=\"https://math.stackexchange.com/questions/1423720/is-there-any-proof-that-the-riemann-zeta-function-is-not-elementary\">here</a> with five (commented and linked) references. </p>\n\n<p>On the other hand, this function is linked with many other transcendental special functions like polylogarithms which satisfy <a href=\"https://arxiv.org/abs/1101.4497v1\" rel=\"nofollow noreferrer\">Fuschian non commutative differential equations</a>. </p>\n" }, { "answer_id": 260208, "score": 21, "text": "<p>The Riemann zeta function is \"hypertranscendental\" in the sense shown <a href=\"https://en.wikipedia.org/wiki/Hypertranscendental_function\">HERE</a></p>\n\n<p>It is not the solution $y(x)$ of a differential equation of the form\n$$\nF(x,y,y',y'',\\dots,y^{(n)})=0\n$$\nwhere $F$ is a polynomial (with constant coefficients).</p>\n" } ]
Question: Could the Riemann zeta function be a solution for a known differential equation? Answer 1: <p>When posed properly, a long-standing open problem, but in the form you ask:</p> <p><em>Robert A. Van Gorder</em>, MR 3276353 <a href="http://dx.doi.org/10.1016/j.jnt.2014.08.013"><strong>Does the Riemann zeta function satisfy a differential equation?</strong></a>, <em>J. Number Theory</em> <strong>147</strong> (2015), 778--788.</p> Answer 2: <p>The fact that $\zeta$ satisfy no algebraic differential equation is due to its famous relation with the Gamma function which was proved by Hölder not to satisfy such an equation.</p> <p>Detailed answer can be found <a href="https://math.stackexchange.com/questions/1423720/is-there-any-proof-that-the-riemann-zeta-function-is-not-elementary">here</a> with five (commented and linked) references. </p> <p>On the other hand, this function is linked with many other transcendental special functions like polylogarithms which satisfy <a href="https://arxiv.org/abs/1101.4497v1" rel="nofollow noreferrer">Fuschian non commutative differential equations</a>. </p> Answer 3: <p>The Riemann zeta function is "hypertranscendental" in the sense shown <a href="https://en.wikipedia.org/wiki/Hypertranscendental_function">HERE</a></p> <p>It is not the solution $y(x)$ of a differential equation of the form $$ F(x,y,y',y'',\dots,y^{(n)})=0 $$ where $F$ is a polynomial (with constant coefficients).</p> Answer 4: <p>When posed properly, a long-standing open problem, but in the form you ask:</p> <p><em>Robert A. Van Gorder</em>, MR 3276353 <a href="http://dx.doi.org/10.1016/j.jnt.2014.08.013"><strong>Does the Riemann zeta function satisfy a differential equation?</strong></a>, <em>J. Number Theory</em> <strong>147</strong> (2015), 778--788.</p> Answer 5: <p>The fact that $\zeta$ satisfy no algebraic differential equation is due to its famous relation with the Gamma function which was proved by Hölder not to satisfy such an equation.</p> <p>Detailed answer can be found <a href="https://math.stackexchange.com/questions/1423720/is-there-any-proof-that-the-riemann-zeta-function-is-not-elementary">here</a> with five (commented and linked) references. </p> <p>On the other hand, this function is linked with many other transcendental special functions like polylogarithms which satisfy <a href="https://arxiv.org/abs/1101.4497v1" rel="nofollow noreferrer">Fuschian non commutative differential equations</a>. </p> Answer 6: <p>The Riemann zeta function is "hypertranscendental" in the sense shown <a href="https://en.wikipedia.org/wiki/Hypertranscendental_function">HERE</a></p> <p>It is not the solution $y(x)$ of a differential equation of the form $$ F(x,y,y',y'',\dots,y^{(n)})=0 $$ where $F$ is a polynomial (with constant coefficients).</p> Answer 7: <p>When posed properly, a long-standing open problem, but in the form you ask:</p> <p><em>Robert A. Van Gorder</em>, MR 3276353 <a href="http://dx.doi.org/10.1016/j.jnt.2014.08.013"><strong>Does the Riemann zeta function satisfy a differential equation?</strong></a>, <em>J. Number Theory</em> <strong>147</strong> (2015), 778--788.</p> Answer 8: <p>The fact that $\zeta$ satisfy no algebraic differential equation is due to its famous relation with the Gamma function which was proved by Hölder not to satisfy such an equation.</p> <p>Detailed answer can be found <a href="https://math.stackexchange.com/questions/1423720/is-there-any-proof-that-the-riemann-zeta-function-is-not-elementary">here</a> with five (commented and linked) references. </p> <p>On the other hand, this function is linked with many other transcendental special functions like polylogarithms which satisfy <a href="https://arxiv.org/abs/1101.4497v1" rel="nofollow noreferrer">Fuschian non commutative differential equations</a>. </p> Answer 9: <p>The Riemann zeta function is "hypertranscendental" in the sense shown <a href="https://en.wikipedia.org/wiki/Hypertranscendental_function">HERE</a></p> <p>It is not the solution $y(x)$ of a differential equation of the form $$ F(x,y,y',y'',\dots,y^{(n)})=0 $$ where $F$ is a polynomial (with constant coefficients).</p> Answer 10: <p>When posed properly, a long-standing open problem, but in the form you ask:</p> <p><em>Robert A. Van Gorder</em>, MR 3276353 <a href="http://dx.doi.org/10.1016/j.jnt.2014.08.013"><strong>Does the Riemann zeta function satisfy a differential equation?</strong></a>, <em>J. Number Theory</em> <strong>147</strong> (2015), 778--788.</p> Answer 11: <p>The fact that $\zeta$ satisfy no algebraic differential equation is due to its famous relation with the Gamma function which was proved by Hölder not to satisfy such an equation.</p> <p>Detailed answer can be found <a href="https://math.stackexchange.com/questions/1423720/is-there-any-proof-that-the-riemann-zeta-function-is-not-elementary">here</a> with five (commented and linked) references. </p> <p>On the other hand, this function is linked with many other transcendental special functions like polylogarithms which satisfy <a href="https://arxiv.org/abs/1101.4497v1" rel="nofollow noreferrer">Fuschian non commutative differential equations</a>. </p> Answer 12: <p>The Riemann zeta function is "hypertranscendental" in the sense shown <a href="https://en.wikipedia.org/wiki/Hypertranscendental_function">HERE</a></p> <p>It is not the solution $y(x)$ of a differential equation of the form $$ F(x,y,y',y'',\dots,y^{(n)})=0 $$ where $F$ is a polynomial (with constant coefficients).</p>
Is the Hurewicz theorem ever used to compute abelianizations?
260,584
32
https://mathoverflow.net/questions/260584/is-the-hurewicz-theorem-ever-used-to-compute-abelianizations
<p>The Hurewicz theorem tells us that if $X$ is a path-connected space then $H_1(X, \, \mathbb{Z})$ is isomorphic to the abelianisation of $\pi_1(X)$. This gives a potential method for computing the abelianisation of a (sufficiently nice) group $G$: realise it as the fundamental group of a space $X$ and then compute $H_1(X, \, \mathbb{Z})$ by your favourite means. </p> <blockquote> <p>Is this method ever used in practice? Are there nice examples of abelianisations which are easily (best?) computed in this way?</p> </blockquote>
[ { "answer_id": 260601, "score": 38, "text": "<p>Yes, the fundamental group of the Hawaiian earring $\\pi_1(\\mathbb{H},b_0)$ is an important group which is sometimes called the free sigma product $\\#_{\\mathbb{N}}\\mathbb{Z}$. Its is often defined in purely algebraic terms as a group of \"transfinite words\" in countably many letters. In many ways this group behaves like the non-abelian version of the Specker group $\\prod_{\\mathbb{N}}\\mathbb{Z}$ and it is the key to the homotopy classification of 1-dimensional Peano continua. However, it's abelianization is not $\\prod_{\\mathbb{N}}\\mathbb{Z}$; it is a good deal more complicated. The abelianization was first described by Eda and Kawamura in the following paper:</p>\n\n<p><em>The Singular Homology Group of the Hawaiian Earring</em>, Journal of the London Mathematical Society, 62 no. 1 (2000) 305–310.</p>\n\n<p>The proof explicitly uses the Hawaiian earring and the authors mention in Remark 2.7 that </p>\n\n<blockquote>\n <p>Though the question itself is formulated algebraically, we have not succeeded in finding a purely algebraic proof of it.</p>\n</blockquote>\n\n<p>One way to represent the isomorphism class of this abelianization is as $$Ab(\\#_{\\mathbb{N}}\\mathbb{Z})\\cong \\prod_{\\mathbb{N}}\\mathbb{Z}\\oplus \\prod_{\\mathbb{N}}\\mathbb{Z}/\\bigoplus_{\\mathbb{N}}\\mathbb{Z}$$</p>\n\n<p>Similar methods have been used to identify abelienizations of fundamental groups of other 1-dimensional spaces, e.g. the Menger curve.</p>\n" }, { "answer_id": 260604, "score": 18, "text": "<p>A usually hard problem in Algebraic Geometry is computing the fundamental group of the complement $U=\\mathbb{P}^n - V$, where $V$ is a reduced hypersurface. </p>\n\n<p>At least in principle, the computation of $\\pi_1(U)$ can be carried out by using Seifert-Van Kampen theorem; the tricky part is that the relations in the presentation depend not only on the singularities of $V$, but also on their mutual position. </p>\n\n<p>A classical example due to Zariski is when $V \\subset \\mathbb{P}^2$ is a curve of degree $6$ having six ordinary cusps and no other singularities. Then $\\pi_1(U)$ is the free product $\\mathbb{Z}/2 \\mathbb{Z} * \\mathbb{Z}/3 \\mathbb{Z}$ if the six cusps lie on the same conic, and the direct product $\\mathbb{Z}/ 2 \\mathbb{Z} \\times \\mathbb{Z}/3 \\mathbb{Z}$ otherwise. </p>\n\n<p>However, by using Lefschetz duality we can compute in a straightforward way $H_1(U, \\, \\mathbb{Z})$ for all hypersurfaces, so that the abelianization of $\\pi_1(U)$ is always known even if a presentation for $\\pi_1(U)$ is not available. In fact, there is the following result that can be found in A. Dimca's book <em>Singularities and topology of hypersurfaces</em>, Chapter 4.</p>\n\n<blockquote>\n <p><strong>Proposition.</strong> Assume that the hypersurface $V \\subset \\mathbb{P}^n$ has $k$ irreducible components $V_1, \\ldots, V_k$ of degrees $d_1, \\ldots, d_k$. Let $d$ be the greatest common divisor of the integers $d_1, \\ldots, d_k$. Then $$H_1(U, \\, \\mathbb{Z}) = \\mathbb{Z}^{k-1} \\oplus \\mathbb{Z}/ d \\mathbb{Z}.$$ </p>\n</blockquote>\n\n<p>In particular, $H_1(U)$ does not depend on the singularities of $V$, but only on the number and degrees of its irreducible components. As a consequence, we recover the fact that in both cases of Zariski's example quoted above (where $k=1, \\, d_1=6$) the abelianization of $\\pi_1(U)$ is the cyclic group of order $6$.</p>\n" }, { "answer_id": 260616, "score": 13, "text": "<p>The mapping class group of a smooth manifold $M$ is the group of all its self diffeomorphisms up to isotopy, i.e. $\\pi_0(\\operatorname{Diff}(M))\\cong \\pi_1(B\\operatorname{Diff}(M))$.</p>\n\n<p>A large portion of geometric topology is concerned with gaining a better understanding of the homotopy type of $B\\operatorname{Diff}(M)$, in particular of its cohomology as the latter is the ring of characteristic class of smooth $M$ bundles.</p>\n\n<p>In the last 15 years, initiated by Madsen and Weiss' solution of the Mumford conjecture, the understanding of the homology of $B\\operatorname{Diff}(M)$ has fundamentally improved which led, among others, to calculations of $H_1(B\\operatorname{Diff}(M))\\cong \\pi_0(\\operatorname{Diff}(M))^{ab}$ for some $M$, e.g. by Galatius and Randal-Williams in the following paper.</p>\n\n<p><em>Abelian quotients of mapping class groups of highly connected manifolds</em>, Mathematische Annalen\nJune 2016, Volume 365, Issue 1, pp 857–879</p>\n" }, { "answer_id": 260601, "score": 38, "text": "<p>Yes, the fundamental group of the Hawaiian earring $\\pi_1(\\mathbb{H},b_0)$ is an important group which is sometimes called the free sigma product $\\#_{\\mathbb{N}}\\mathbb{Z}$. Its is often defined in purely algebraic terms as a group of \"transfinite words\" in countably many letters. In many ways this group behaves like the non-abelian version of the Specker group $\\prod_{\\mathbb{N}}\\mathbb{Z}$ and it is the key to the homotopy classification of 1-dimensional Peano continua. However, it's abelianization is not $\\prod_{\\mathbb{N}}\\mathbb{Z}$; it is a good deal more complicated. The abelianization was first described by Eda and Kawamura in the following paper:</p>\n\n<p><em>The Singular Homology Group of the Hawaiian Earring</em>, Journal of the London Mathematical Society, 62 no. 1 (2000) 305–310.</p>\n\n<p>The proof explicitly uses the Hawaiian earring and the authors mention in Remark 2.7 that </p>\n\n<blockquote>\n <p>Though the question itself is formulated algebraically, we have not succeeded in finding a purely algebraic proof of it.</p>\n</blockquote>\n\n<p>One way to represent the isomorphism class of this abelianization is as $$Ab(\\#_{\\mathbb{N}}\\mathbb{Z})\\cong \\prod_{\\mathbb{N}}\\mathbb{Z}\\oplus \\prod_{\\mathbb{N}}\\mathbb{Z}/\\bigoplus_{\\mathbb{N}}\\mathbb{Z}$$</p>\n\n<p>Similar methods have been used to identify abelienizations of fundamental groups of other 1-dimensional spaces, e.g. the Menger curve.</p>\n" }, { "answer_id": 260604, "score": 18, "text": "<p>A usually hard problem in Algebraic Geometry is computing the fundamental group of the complement $U=\\mathbb{P}^n - V$, where $V$ is a reduced hypersurface. </p>\n\n<p>At least in principle, the computation of $\\pi_1(U)$ can be carried out by using Seifert-Van Kampen theorem; the tricky part is that the relations in the presentation depend not only on the singularities of $V$, but also on their mutual position. </p>\n\n<p>A classical example due to Zariski is when $V \\subset \\mathbb{P}^2$ is a curve of degree $6$ having six ordinary cusps and no other singularities. Then $\\pi_1(U)$ is the free product $\\mathbb{Z}/2 \\mathbb{Z} * \\mathbb{Z}/3 \\mathbb{Z}$ if the six cusps lie on the same conic, and the direct product $\\mathbb{Z}/ 2 \\mathbb{Z} \\times \\mathbb{Z}/3 \\mathbb{Z}$ otherwise. </p>\n\n<p>However, by using Lefschetz duality we can compute in a straightforward way $H_1(U, \\, \\mathbb{Z})$ for all hypersurfaces, so that the abelianization of $\\pi_1(U)$ is always known even if a presentation for $\\pi_1(U)$ is not available. In fact, there is the following result that can be found in A. Dimca's book <em>Singularities and topology of hypersurfaces</em>, Chapter 4.</p>\n\n<blockquote>\n <p><strong>Proposition.</strong> Assume that the hypersurface $V \\subset \\mathbb{P}^n$ has $k$ irreducible components $V_1, \\ldots, V_k$ of degrees $d_1, \\ldots, d_k$. Let $d$ be the greatest common divisor of the integers $d_1, \\ldots, d_k$. Then $$H_1(U, \\, \\mathbb{Z}) = \\mathbb{Z}^{k-1} \\oplus \\mathbb{Z}/ d \\mathbb{Z}.$$ </p>\n</blockquote>\n\n<p>In particular, $H_1(U)$ does not depend on the singularities of $V$, but only on the number and degrees of its irreducible components. As a consequence, we recover the fact that in both cases of Zariski's example quoted above (where $k=1, \\, d_1=6$) the abelianization of $\\pi_1(U)$ is the cyclic group of order $6$.</p>\n" }, { "answer_id": 260616, "score": 13, "text": "<p>The mapping class group of a smooth manifold $M$ is the group of all its self diffeomorphisms up to isotopy, i.e. $\\pi_0(\\operatorname{Diff}(M))\\cong \\pi_1(B\\operatorname{Diff}(M))$.</p>\n\n<p>A large portion of geometric topology is concerned with gaining a better understanding of the homotopy type of $B\\operatorname{Diff}(M)$, in particular of its cohomology as the latter is the ring of characteristic class of smooth $M$ bundles.</p>\n\n<p>In the last 15 years, initiated by Madsen and Weiss' solution of the Mumford conjecture, the understanding of the homology of $B\\operatorname{Diff}(M)$ has fundamentally improved which led, among others, to calculations of $H_1(B\\operatorname{Diff}(M))\\cong \\pi_0(\\operatorname{Diff}(M))^{ab}$ for some $M$, e.g. by Galatius and Randal-Williams in the following paper.</p>\n\n<p><em>Abelian quotients of mapping class groups of highly connected manifolds</em>, Mathematische Annalen\nJune 2016, Volume 365, Issue 1, pp 857–879</p>\n" }, { "answer_id": 260601, "score": 38, "text": "<p>Yes, the fundamental group of the Hawaiian earring $\\pi_1(\\mathbb{H},b_0)$ is an important group which is sometimes called the free sigma product $\\#_{\\mathbb{N}}\\mathbb{Z}$. Its is often defined in purely algebraic terms as a group of \"transfinite words\" in countably many letters. In many ways this group behaves like the non-abelian version of the Specker group $\\prod_{\\mathbb{N}}\\mathbb{Z}$ and it is the key to the homotopy classification of 1-dimensional Peano continua. However, it's abelianization is not $\\prod_{\\mathbb{N}}\\mathbb{Z}$; it is a good deal more complicated. The abelianization was first described by Eda and Kawamura in the following paper:</p>\n\n<p><em>The Singular Homology Group of the Hawaiian Earring</em>, Journal of the London Mathematical Society, 62 no. 1 (2000) 305–310.</p>\n\n<p>The proof explicitly uses the Hawaiian earring and the authors mention in Remark 2.7 that </p>\n\n<blockquote>\n <p>Though the question itself is formulated algebraically, we have not succeeded in finding a purely algebraic proof of it.</p>\n</blockquote>\n\n<p>One way to represent the isomorphism class of this abelianization is as $$Ab(\\#_{\\mathbb{N}}\\mathbb{Z})\\cong \\prod_{\\mathbb{N}}\\mathbb{Z}\\oplus \\prod_{\\mathbb{N}}\\mathbb{Z}/\\bigoplus_{\\mathbb{N}}\\mathbb{Z}$$</p>\n\n<p>Similar methods have been used to identify abelienizations of fundamental groups of other 1-dimensional spaces, e.g. the Menger curve.</p>\n" }, { "answer_id": 260604, "score": 18, "text": "<p>A usually hard problem in Algebraic Geometry is computing the fundamental group of the complement $U=\\mathbb{P}^n - V$, where $V$ is a reduced hypersurface. </p>\n\n<p>At least in principle, the computation of $\\pi_1(U)$ can be carried out by using Seifert-Van Kampen theorem; the tricky part is that the relations in the presentation depend not only on the singularities of $V$, but also on their mutual position. </p>\n\n<p>A classical example due to Zariski is when $V \\subset \\mathbb{P}^2$ is a curve of degree $6$ having six ordinary cusps and no other singularities. Then $\\pi_1(U)$ is the free product $\\mathbb{Z}/2 \\mathbb{Z} * \\mathbb{Z}/3 \\mathbb{Z}$ if the six cusps lie on the same conic, and the direct product $\\mathbb{Z}/ 2 \\mathbb{Z} \\times \\mathbb{Z}/3 \\mathbb{Z}$ otherwise. </p>\n\n<p>However, by using Lefschetz duality we can compute in a straightforward way $H_1(U, \\, \\mathbb{Z})$ for all hypersurfaces, so that the abelianization of $\\pi_1(U)$ is always known even if a presentation for $\\pi_1(U)$ is not available. In fact, there is the following result that can be found in A. Dimca's book <em>Singularities and topology of hypersurfaces</em>, Chapter 4.</p>\n\n<blockquote>\n <p><strong>Proposition.</strong> Assume that the hypersurface $V \\subset \\mathbb{P}^n$ has $k$ irreducible components $V_1, \\ldots, V_k$ of degrees $d_1, \\ldots, d_k$. Let $d$ be the greatest common divisor of the integers $d_1, \\ldots, d_k$. Then $$H_1(U, \\, \\mathbb{Z}) = \\mathbb{Z}^{k-1} \\oplus \\mathbb{Z}/ d \\mathbb{Z}.$$ </p>\n</blockquote>\n\n<p>In particular, $H_1(U)$ does not depend on the singularities of $V$, but only on the number and degrees of its irreducible components. As a consequence, we recover the fact that in both cases of Zariski's example quoted above (where $k=1, \\, d_1=6$) the abelianization of $\\pi_1(U)$ is the cyclic group of order $6$.</p>\n" }, { "answer_id": 260616, "score": 13, "text": "<p>The mapping class group of a smooth manifold $M$ is the group of all its self diffeomorphisms up to isotopy, i.e. $\\pi_0(\\operatorname{Diff}(M))\\cong \\pi_1(B\\operatorname{Diff}(M))$.</p>\n\n<p>A large portion of geometric topology is concerned with gaining a better understanding of the homotopy type of $B\\operatorname{Diff}(M)$, in particular of its cohomology as the latter is the ring of characteristic class of smooth $M$ bundles.</p>\n\n<p>In the last 15 years, initiated by Madsen and Weiss' solution of the Mumford conjecture, the understanding of the homology of $B\\operatorname{Diff}(M)$ has fundamentally improved which led, among others, to calculations of $H_1(B\\operatorname{Diff}(M))\\cong \\pi_0(\\operatorname{Diff}(M))^{ab}$ for some $M$, e.g. by Galatius and Randal-Williams in the following paper.</p>\n\n<p><em>Abelian quotients of mapping class groups of highly connected manifolds</em>, Mathematische Annalen\nJune 2016, Volume 365, Issue 1, pp 857–879</p>\n" }, { "answer_id": 260601, "score": 38, "text": "<p>Yes, the fundamental group of the Hawaiian earring $\\pi_1(\\mathbb{H},b_0)$ is an important group which is sometimes called the free sigma product $\\#_{\\mathbb{N}}\\mathbb{Z}$. Its is often defined in purely algebraic terms as a group of \"transfinite words\" in countably many letters. In many ways this group behaves like the non-abelian version of the Specker group $\\prod_{\\mathbb{N}}\\mathbb{Z}$ and it is the key to the homotopy classification of 1-dimensional Peano continua. However, it's abelianization is not $\\prod_{\\mathbb{N}}\\mathbb{Z}$; it is a good deal more complicated. The abelianization was first described by Eda and Kawamura in the following paper:</p>\n\n<p><em>The Singular Homology Group of the Hawaiian Earring</em>, Journal of the London Mathematical Society, 62 no. 1 (2000) 305–310.</p>\n\n<p>The proof explicitly uses the Hawaiian earring and the authors mention in Remark 2.7 that </p>\n\n<blockquote>\n <p>Though the question itself is formulated algebraically, we have not succeeded in finding a purely algebraic proof of it.</p>\n</blockquote>\n\n<p>One way to represent the isomorphism class of this abelianization is as $$Ab(\\#_{\\mathbb{N}}\\mathbb{Z})\\cong \\prod_{\\mathbb{N}}\\mathbb{Z}\\oplus \\prod_{\\mathbb{N}}\\mathbb{Z}/\\bigoplus_{\\mathbb{N}}\\mathbb{Z}$$</p>\n\n<p>Similar methods have been used to identify abelienizations of fundamental groups of other 1-dimensional spaces, e.g. the Menger curve.</p>\n" }, { "answer_id": 260604, "score": 18, "text": "<p>A usually hard problem in Algebraic Geometry is computing the fundamental group of the complement $U=\\mathbb{P}^n - V$, where $V$ is a reduced hypersurface. </p>\n\n<p>At least in principle, the computation of $\\pi_1(U)$ can be carried out by using Seifert-Van Kampen theorem; the tricky part is that the relations in the presentation depend not only on the singularities of $V$, but also on their mutual position. </p>\n\n<p>A classical example due to Zariski is when $V \\subset \\mathbb{P}^2$ is a curve of degree $6$ having six ordinary cusps and no other singularities. Then $\\pi_1(U)$ is the free product $\\mathbb{Z}/2 \\mathbb{Z} * \\mathbb{Z}/3 \\mathbb{Z}$ if the six cusps lie on the same conic, and the direct product $\\mathbb{Z}/ 2 \\mathbb{Z} \\times \\mathbb{Z}/3 \\mathbb{Z}$ otherwise. </p>\n\n<p>However, by using Lefschetz duality we can compute in a straightforward way $H_1(U, \\, \\mathbb{Z})$ for all hypersurfaces, so that the abelianization of $\\pi_1(U)$ is always known even if a presentation for $\\pi_1(U)$ is not available. In fact, there is the following result that can be found in A. Dimca's book <em>Singularities and topology of hypersurfaces</em>, Chapter 4.</p>\n\n<blockquote>\n <p><strong>Proposition.</strong> Assume that the hypersurface $V \\subset \\mathbb{P}^n$ has $k$ irreducible components $V_1, \\ldots, V_k$ of degrees $d_1, \\ldots, d_k$. Let $d$ be the greatest common divisor of the integers $d_1, \\ldots, d_k$. Then $$H_1(U, \\, \\mathbb{Z}) = \\mathbb{Z}^{k-1} \\oplus \\mathbb{Z}/ d \\mathbb{Z}.$$ </p>\n</blockquote>\n\n<p>In particular, $H_1(U)$ does not depend on the singularities of $V$, but only on the number and degrees of its irreducible components. As a consequence, we recover the fact that in both cases of Zariski's example quoted above (where $k=1, \\, d_1=6$) the abelianization of $\\pi_1(U)$ is the cyclic group of order $6$.</p>\n" }, { "answer_id": 260616, "score": 13, "text": "<p>The mapping class group of a smooth manifold $M$ is the group of all its self diffeomorphisms up to isotopy, i.e. $\\pi_0(\\operatorname{Diff}(M))\\cong \\pi_1(B\\operatorname{Diff}(M))$.</p>\n\n<p>A large portion of geometric topology is concerned with gaining a better understanding of the homotopy type of $B\\operatorname{Diff}(M)$, in particular of its cohomology as the latter is the ring of characteristic class of smooth $M$ bundles.</p>\n\n<p>In the last 15 years, initiated by Madsen and Weiss' solution of the Mumford conjecture, the understanding of the homology of $B\\operatorname{Diff}(M)$ has fundamentally improved which led, among others, to calculations of $H_1(B\\operatorname{Diff}(M))\\cong \\pi_0(\\operatorname{Diff}(M))^{ab}$ for some $M$, e.g. by Galatius and Randal-Williams in the following paper.</p>\n\n<p><em>Abelian quotients of mapping class groups of highly connected manifolds</em>, Mathematische Annalen\nJune 2016, Volume 365, Issue 1, pp 857–879</p>\n" } ]
Question: Is the Hurewicz theorem ever used to compute abelianizations? Answer 1: <p>Yes, the fundamental group of the Hawaiian earring $\pi_1(\mathbb{H},b_0)$ is an important group which is sometimes called the free sigma product $\#_{\mathbb{N}}\mathbb{Z}$. Its is often defined in purely algebraic terms as a group of "transfinite words" in countably many letters. In many ways this group behaves like the non-abelian version of the Specker group $\prod_{\mathbb{N}}\mathbb{Z}$ and it is the key to the homotopy classification of 1-dimensional Peano continua. However, it's abelianization is not $\prod_{\mathbb{N}}\mathbb{Z}$; it is a good deal more complicated. The abelianization was first described by Eda and Kawamura in the following paper:</p> <p><em>The Singular Homology Group of the Hawaiian Earring</em>, Journal of the London Mathematical Society, 62 no. 1 (2000) 305–310.</p> <p>The proof explicitly uses the Hawaiian earring and the authors mention in Remark 2.7 that </p> <blockquote> <p>Though the question itself is formulated algebraically, we have not succeeded in finding a purely algebraic proof of it.</p> </blockquote> <p>One way to represent the isomorphism class of this abelianization is as $$Ab(\#_{\mathbb{N}}\mathbb{Z})\cong \prod_{\mathbb{N}}\mathbb{Z}\oplus \prod_{\mathbb{N}}\mathbb{Z}/\bigoplus_{\mathbb{N}}\mathbb{Z}$$</p> <p>Similar methods have been used to identify abelienizations of fundamental groups of other 1-dimensional spaces, e.g. the Menger curve.</p> Answer 2: <p>A usually hard problem in Algebraic Geometry is computing the fundamental group of the complement $U=\mathbb{P}^n - V$, where $V$ is a reduced hypersurface. </p> <p>At least in principle, the computation of $\pi_1(U)$ can be carried out by using Seifert-Van Kampen theorem; the tricky part is that the relations in the presentation depend not only on the singularities of $V$, but also on their mutual position. </p> <p>A classical example due to Zariski is when $V \subset \mathbb{P}^2$ is a curve of degree $6$ having six ordinary cusps and no other singularities. Then $\pi_1(U)$ is the free product $\mathbb{Z}/2 \mathbb{Z} * \mathbb{Z}/3 \mathbb{Z}$ if the six cusps lie on the same conic, and the direct product $\mathbb{Z}/ 2 \mathbb{Z} \times \mathbb{Z}/3 \mathbb{Z}$ otherwise. </p> <p>However, by using Lefschetz duality we can compute in a straightforward way $H_1(U, \, \mathbb{Z})$ for all hypersurfaces, so that the abelianization of $\pi_1(U)$ is always known even if a presentation for $\pi_1(U)$ is not available. In fact, there is the following result that can be found in A. Dimca's book <em>Singularities and topology of hypersurfaces</em>, Chapter 4.</p> <blockquote> <p><strong>Proposition.</strong> Assume that the hypersurface $V \subset \mathbb{P}^n$ has $k$ irreducible components $V_1, \ldots, V_k$ of degrees $d_1, \ldots, d_k$. Let $d$ be the greatest common divisor of the integers $d_1, \ldots, d_k$. Then $$H_1(U, \, \mathbb{Z}) = \mathbb{Z}^{k-1} \oplus \mathbb{Z}/ d \mathbb{Z}.$$ </p> </blockquote> <p>In particular, $H_1(U)$ does not depend on the singularities of $V$, but only on the number and degrees of its irreducible components. As a consequence, we recover the fact that in both cases of Zariski's example quoted above (where $k=1, \, d_1=6$) the abelianization of $\pi_1(U)$ is the cyclic group of order $6$.</p> Answer 3: <p>The mapping class group of a smooth manifold $M$ is the group of all its self diffeomorphisms up to isotopy, i.e. $\pi_0(\operatorname{Diff}(M))\cong \pi_1(B\operatorname{Diff}(M))$.</p> <p>A large portion of geometric topology is concerned with gaining a better understanding of the homotopy type of $B\operatorname{Diff}(M)$, in particular of its cohomology as the latter is the ring of characteristic class of smooth $M$ bundles.</p> <p>In the last 15 years, initiated by Madsen and Weiss' solution of the Mumford conjecture, the understanding of the homology of $B\operatorname{Diff}(M)$ has fundamentally improved which led, among others, to calculations of $H_1(B\operatorname{Diff}(M))\cong \pi_0(\operatorname{Diff}(M))^{ab}$ for some $M$, e.g. by Galatius and Randal-Williams in the following paper.</p> <p><em>Abelian quotients of mapping class groups of highly connected manifolds</em>, Mathematische Annalen June 2016, Volume 365, Issue 1, pp 857–879</p> Answer 4: <p>Yes, the fundamental group of the Hawaiian earring $\pi_1(\mathbb{H},b_0)$ is an important group which is sometimes called the free sigma product $\#_{\mathbb{N}}\mathbb{Z}$. Its is often defined in purely algebraic terms as a group of "transfinite words" in countably many letters. In many ways this group behaves like the non-abelian version of the Specker group $\prod_{\mathbb{N}}\mathbb{Z}$ and it is the key to the homotopy classification of 1-dimensional Peano continua. However, it's abelianization is not $\prod_{\mathbb{N}}\mathbb{Z}$; it is a good deal more complicated. The abelianization was first described by Eda and Kawamura in the following paper:</p> <p><em>The Singular Homology Group of the Hawaiian Earring</em>, Journal of the London Mathematical Society, 62 no. 1 (2000) 305–310.</p> <p>The proof explicitly uses the Hawaiian earring and the authors mention in Remark 2.7 that </p> <blockquote> <p>Though the question itself is formulated algebraically, we have not succeeded in finding a purely algebraic proof of it.</p> </blockquote> <p>One way to represent the isomorphism class of this abelianization is as $$Ab(\#_{\mathbb{N}}\mathbb{Z})\cong \prod_{\mathbb{N}}\mathbb{Z}\oplus \prod_{\mathbb{N}}\mathbb{Z}/\bigoplus_{\mathbb{N}}\mathbb{Z}$$</p> <p>Similar methods have been used to identify abelienizations of fundamental groups of other 1-dimensional spaces, e.g. the Menger curve.</p> Answer 5: <p>A usually hard problem in Algebraic Geometry is computing the fundamental group of the complement $U=\mathbb{P}^n - V$, where $V$ is a reduced hypersurface. </p> <p>At least in principle, the computation of $\pi_1(U)$ can be carried out by using Seifert-Van Kampen theorem; the tricky part is that the relations in the presentation depend not only on the singularities of $V$, but also on their mutual position. </p> <p>A classical example due to Zariski is when $V \subset \mathbb{P}^2$ is a curve of degree $6$ having six ordinary cusps and no other singularities. Then $\pi_1(U)$ is the free product $\mathbb{Z}/2 \mathbb{Z} * \mathbb{Z}/3 \mathbb{Z}$ if the six cusps lie on the same conic, and the direct product $\mathbb{Z}/ 2 \mathbb{Z} \times \mathbb{Z}/3 \mathbb{Z}$ otherwise. </p> <p>However, by using Lefschetz duality we can compute in a straightforward way $H_1(U, \, \mathbb{Z})$ for all hypersurfaces, so that the abelianization of $\pi_1(U)$ is always known even if a presentation for $\pi_1(U)$ is not available. In fact, there is the following result that can be found in A. Dimca's book <em>Singularities and topology of hypersurfaces</em>, Chapter 4.</p> <blockquote> <p><strong>Proposition.</strong> Assume that the hypersurface $V \subset \mathbb{P}^n$ has $k$ irreducible components $V_1, \ldots, V_k$ of degrees $d_1, \ldots, d_k$. Let $d$ be the greatest common divisor of the integers $d_1, \ldots, d_k$. Then $$H_1(U, \, \mathbb{Z}) = \mathbb{Z}^{k-1} \oplus \mathbb{Z}/ d \mathbb{Z}.$$ </p> </blockquote> <p>In particular, $H_1(U)$ does not depend on the singularities of $V$, but only on the number and degrees of its irreducible components. As a consequence, we recover the fact that in both cases of Zariski's example quoted above (where $k=1, \, d_1=6$) the abelianization of $\pi_1(U)$ is the cyclic group of order $6$.</p> Answer 6: <p>The mapping class group of a smooth manifold $M$ is the group of all its self diffeomorphisms up to isotopy, i.e. $\pi_0(\operatorname{Diff}(M))\cong \pi_1(B\operatorname{Diff}(M))$.</p> <p>A large portion of geometric topology is concerned with gaining a better understanding of the homotopy type of $B\operatorname{Diff}(M)$, in particular of its cohomology as the latter is the ring of characteristic class of smooth $M$ bundles.</p> <p>In the last 15 years, initiated by Madsen and Weiss' solution of the Mumford conjecture, the understanding of the homology of $B\operatorname{Diff}(M)$ has fundamentally improved which led, among others, to calculations of $H_1(B\operatorname{Diff}(M))\cong \pi_0(\operatorname{Diff}(M))^{ab}$ for some $M$, e.g. by Galatius and Randal-Williams in the following paper.</p> <p><em>Abelian quotients of mapping class groups of highly connected manifolds</em>, Mathematische Annalen June 2016, Volume 365, Issue 1, pp 857–879</p> Answer 7: <p>Yes, the fundamental group of the Hawaiian earring $\pi_1(\mathbb{H},b_0)$ is an important group which is sometimes called the free sigma product $\#_{\mathbb{N}}\mathbb{Z}$. Its is often defined in purely algebraic terms as a group of "transfinite words" in countably many letters. In many ways this group behaves like the non-abelian version of the Specker group $\prod_{\mathbb{N}}\mathbb{Z}$ and it is the key to the homotopy classification of 1-dimensional Peano continua. However, it's abelianization is not $\prod_{\mathbb{N}}\mathbb{Z}$; it is a good deal more complicated. The abelianization was first described by Eda and Kawamura in the following paper:</p> <p><em>The Singular Homology Group of the Hawaiian Earring</em>, Journal of the London Mathematical Society, 62 no. 1 (2000) 305–310.</p> <p>The proof explicitly uses the Hawaiian earring and the authors mention in Remark 2.7 that </p> <blockquote> <p>Though the question itself is formulated algebraically, we have not succeeded in finding a purely algebraic proof of it.</p> </blockquote> <p>One way to represent the isomorphism class of this abelianization is as $$Ab(\#_{\mathbb{N}}\mathbb{Z})\cong \prod_{\mathbb{N}}\mathbb{Z}\oplus \prod_{\mathbb{N}}\mathbb{Z}/\bigoplus_{\mathbb{N}}\mathbb{Z}$$</p> <p>Similar methods have been used to identify abelienizations of fundamental groups of other 1-dimensional spaces, e.g. the Menger curve.</p> Answer 8: <p>A usually hard problem in Algebraic Geometry is computing the fundamental group of the complement $U=\mathbb{P}^n - V$, where $V$ is a reduced hypersurface. </p> <p>At least in principle, the computation of $\pi_1(U)$ can be carried out by using Seifert-Van Kampen theorem; the tricky part is that the relations in the presentation depend not only on the singularities of $V$, but also on their mutual position. </p> <p>A classical example due to Zariski is when $V \subset \mathbb{P}^2$ is a curve of degree $6$ having six ordinary cusps and no other singularities. Then $\pi_1(U)$ is the free product $\mathbb{Z}/2 \mathbb{Z} * \mathbb{Z}/3 \mathbb{Z}$ if the six cusps lie on the same conic, and the direct product $\mathbb{Z}/ 2 \mathbb{Z} \times \mathbb{Z}/3 \mathbb{Z}$ otherwise. </p> <p>However, by using Lefschetz duality we can compute in a straightforward way $H_1(U, \, \mathbb{Z})$ for all hypersurfaces, so that the abelianization of $\pi_1(U)$ is always known even if a presentation for $\pi_1(U)$ is not available. In fact, there is the following result that can be found in A. Dimca's book <em>Singularities and topology of hypersurfaces</em>, Chapter 4.</p> <blockquote> <p><strong>Proposition.</strong> Assume that the hypersurface $V \subset \mathbb{P}^n$ has $k$ irreducible components $V_1, \ldots, V_k$ of degrees $d_1, \ldots, d_k$. Let $d$ be the greatest common divisor of the integers $d_1, \ldots, d_k$. Then $$H_1(U, \, \mathbb{Z}) = \mathbb{Z}^{k-1} \oplus \mathbb{Z}/ d \mathbb{Z}.$$ </p> </blockquote> <p>In particular, $H_1(U)$ does not depend on the singularities of $V$, but only on the number and degrees of its irreducible components. As a consequence, we recover the fact that in both cases of Zariski's example quoted above (where $k=1, \, d_1=6$) the abelianization of $\pi_1(U)$ is the cyclic group of order $6$.</p> Answer 9: <p>The mapping class group of a smooth manifold $M$ is the group of all its self diffeomorphisms up to isotopy, i.e. $\pi_0(\operatorname{Diff}(M))\cong \pi_1(B\operatorname{Diff}(M))$.</p> <p>A large portion of geometric topology is concerned with gaining a better understanding of the homotopy type of $B\operatorname{Diff}(M)$, in particular of its cohomology as the latter is the ring of characteristic class of smooth $M$ bundles.</p> <p>In the last 15 years, initiated by Madsen and Weiss' solution of the Mumford conjecture, the understanding of the homology of $B\operatorname{Diff}(M)$ has fundamentally improved which led, among others, to calculations of $H_1(B\operatorname{Diff}(M))\cong \pi_0(\operatorname{Diff}(M))^{ab}$ for some $M$, e.g. by Galatius and Randal-Williams in the following paper.</p> <p><em>Abelian quotients of mapping class groups of highly connected manifolds</em>, Mathematische Annalen June 2016, Volume 365, Issue 1, pp 857–879</p> Answer 10: <p>Yes, the fundamental group of the Hawaiian earring $\pi_1(\mathbb{H},b_0)$ is an important group which is sometimes called the free sigma product $\#_{\mathbb{N}}\mathbb{Z}$. Its is often defined in purely algebraic terms as a group of "transfinite words" in countably many letters. In many ways this group behaves like the non-abelian version of the Specker group $\prod_{\mathbb{N}}\mathbb{Z}$ and it is the key to the homotopy classification of 1-dimensional Peano continua. However, it's abelianization is not $\prod_{\mathbb{N}}\mathbb{Z}$; it is a good deal more complicated. The abelianization was first described by Eda and Kawamura in the following paper:</p> <p><em>The Singular Homology Group of the Hawaiian Earring</em>, Journal of the London Mathematical Society, 62 no. 1 (2000) 305–310.</p> <p>The proof explicitly uses the Hawaiian earring and the authors mention in Remark 2.7 that </p> <blockquote> <p>Though the question itself is formulated algebraically, we have not succeeded in finding a purely algebraic proof of it.</p> </blockquote> <p>One way to represent the isomorphism class of this abelianization is as $$Ab(\#_{\mathbb{N}}\mathbb{Z})\cong \prod_{\mathbb{N}}\mathbb{Z}\oplus \prod_{\mathbb{N}}\mathbb{Z}/\bigoplus_{\mathbb{N}}\mathbb{Z}$$</p> <p>Similar methods have been used to identify abelienizations of fundamental groups of other 1-dimensional spaces, e.g. the Menger curve.</p> Answer 11: <p>A usually hard problem in Algebraic Geometry is computing the fundamental group of the complement $U=\mathbb{P}^n - V$, where $V$ is a reduced hypersurface. </p> <p>At least in principle, the computation of $\pi_1(U)$ can be carried out by using Seifert-Van Kampen theorem; the tricky part is that the relations in the presentation depend not only on the singularities of $V$, but also on their mutual position. </p> <p>A classical example due to Zariski is when $V \subset \mathbb{P}^2$ is a curve of degree $6$ having six ordinary cusps and no other singularities. Then $\pi_1(U)$ is the free product $\mathbb{Z}/2 \mathbb{Z} * \mathbb{Z}/3 \mathbb{Z}$ if the six cusps lie on the same conic, and the direct product $\mathbb{Z}/ 2 \mathbb{Z} \times \mathbb{Z}/3 \mathbb{Z}$ otherwise. </p> <p>However, by using Lefschetz duality we can compute in a straightforward way $H_1(U, \, \mathbb{Z})$ for all hypersurfaces, so that the abelianization of $\pi_1(U)$ is always known even if a presentation for $\pi_1(U)$ is not available. In fact, there is the following result that can be found in A. Dimca's book <em>Singularities and topology of hypersurfaces</em>, Chapter 4.</p> <blockquote> <p><strong>Proposition.</strong> Assume that the hypersurface $V \subset \mathbb{P}^n$ has $k$ irreducible components $V_1, \ldots, V_k$ of degrees $d_1, \ldots, d_k$. Let $d$ be the greatest common divisor of the integers $d_1, \ldots, d_k$. Then $$H_1(U, \, \mathbb{Z}) = \mathbb{Z}^{k-1} \oplus \mathbb{Z}/ d \mathbb{Z}.$$ </p> </blockquote> <p>In particular, $H_1(U)$ does not depend on the singularities of $V$, but only on the number and degrees of its irreducible components. As a consequence, we recover the fact that in both cases of Zariski's example quoted above (where $k=1, \, d_1=6$) the abelianization of $\pi_1(U)$ is the cyclic group of order $6$.</p> Answer 12: <p>The mapping class group of a smooth manifold $M$ is the group of all its self diffeomorphisms up to isotopy, i.e. $\pi_0(\operatorname{Diff}(M))\cong \pi_1(B\operatorname{Diff}(M))$.</p> <p>A large portion of geometric topology is concerned with gaining a better understanding of the homotopy type of $B\operatorname{Diff}(M)$, in particular of its cohomology as the latter is the ring of characteristic class of smooth $M$ bundles.</p> <p>In the last 15 years, initiated by Madsen and Weiss' solution of the Mumford conjecture, the understanding of the homology of $B\operatorname{Diff}(M)$ has fundamentally improved which led, among others, to calculations of $H_1(B\operatorname{Diff}(M))\cong \pi_0(\operatorname{Diff}(M))^{ab}$ for some $M$, e.g. by Galatius and Randal-Williams in the following paper.</p> <p><em>Abelian quotients of mapping class groups of highly connected manifolds</em>, Mathematische Annalen June 2016, Volume 365, Issue 1, pp 857–879</p>
Ideas in the elementary proof of the prime number theorem (Selberg / Erdős)
259,698
31
https://mathoverflow.net/questions/259698/ideas-in-the-elementary-proof-of-the-prime-number-theorem-selberg-erd%c5%91s
<p>I'm reading the elementary proof of prime number theorem (Selberg / Erdős, around 1949). </p> <p>One key step is to prove that, with $\vartheta(x) = \sum_{p\leq x} \log p$,</p> <p>$$(1) \qquad\qquad \vartheta(x) \log x+\sum_{p\leq x}(\log p) \vartheta(x/p) = 2x\log x+O(x)$$</p> <ol> <li><p><strong>What's the idea behind this identity?</strong></p> <p>Here is the heuristic argument I understand : if $p$ is a prime number, the average gap before the next prime is $\approx \log p$. Thus if we set a weight $0$ for non-prime numbers, and $\log p$ for prime numbers, then when we sum all these weights for $n = 1 ... x$, we should get $\vartheta(x) = \sum_{p \leq x} \log p \sim x$, and that's equivalent to the PNT. Ok. </p> <p>But where comes the idea from, to add on the left hand side of the identity (1) the part $\sum_{p\leq x}(\log p) \vartheta(x/p)$ ? What's the idea here?</p></li> <li><p><strong>What's the idea to go from (1) to the prime number theorem?</strong></p></li> </ol> <hr> <p><em>Note: the identity (1) can be replaced by <a href="http://projecteuclid.org/download/pdf_1/euclid.pja/1195571353">an equivalent identity</a>, with $\psi(x) = \sum_{p^k\leq x} \log p = \sum_{n \leq x} \Lambda(n)$ (where $\Lambda(n) = \log p$ if $n=p^k$ and $0$ otherwise):</em></p> <p>$$(2) \qquad\qquad \psi(x) \log x+\sum_{n\leq x}\Lambda(n) \psi(x/n) = 2x\log x+O(x)$$</p> <p><em>Another equivalent form is (just in case (2) or (3) might be easier to understand):</em></p> <p>$$(3) \qquad\qquad \sum_{p \leq x} (\log p)^2 + \sum_{pq\leq x} \log p \log q= 2x\log x+O(x)$$</p>
[ { "answer_id": 259719, "score": 52, "text": "<p>The complex-analytic proof of the prime number theorem can help inform the elementary one.</p>\n\n<p>The von Mangoldt function $\\Lambda$ is related to the Riemann zeta function $\\zeta$ by the formula\n$$ \\sum_n \\frac{\\Lambda(n)}{n^s} = -\\frac{\\zeta'(s)}{\\zeta(s)},$$\nat least in the region $\\mathrm{Re}(s) &gt; 1$. The right-hand side has a pole of residue 1 at the simple pole $s=1$ of $\\zeta$, and a pole of residue -1 at each zero $s=\\rho$ of $\\zeta$. Applying Perron's formula carefully, one eventually arrives at the von Mangoldt <a href=\"https://en.wikipedia.org/wiki/Explicit_formulae_(L-function)\" rel=\"noreferrer\">explicit formula</a></p>\n\n<p>$$ \\sum_{n \\leq x} \\Lambda(n) = x - \\sum_\\rho \\frac{x^\\rho}{\\rho} + \\dots$$\nwhere the remaining terms $\\dots$ (as well as the way in which the infinite sum over zeroes is interpreted) will be ignored for this discussion. This formula strongly suggests that zeroes $\\rho$ with real part strictly less than one will eventually only make a lower order contribution to $\\sum_{n \\leq x} \\Lambda(n)$, whereas zeroes $\\rho$ with real part equal to one would likely destroy the prime number theorem. In fact one can pursue this line of reasoning more carefully and show that the prime number theorem is equivalent to the absence of zeroes of zeta on the line $\\mathrm{Re}(s)=1$.</p>\n\n<p>But what if there was a way to make the contribution of the zeroes of zeta smaller than those of the pole, even when said zeroes lie on the line $\\mathrm{Re}(s)=1$? One way to do this is to replace the log-derivative $-\\frac{\\zeta'(s)}{\\zeta(s)}$ by the variant $\\frac{\\zeta''(s)}{\\zeta(s)}$. Assuming for simplicity that all zeroes are simple, this function still has a simple pole at every zero $\\rho$ (with residue $\\frac{\\zeta''(\\rho)}{\\zeta'(\\rho)}$), but now has a <em>double</em> pole at $s=1$ (with residue $2$). As it turns out, this variant also has a Dirichlet series representation\n$$ \\sum_n \\frac{ \\Lambda_2(n) }{n^s} = \\frac{\\zeta''(s)}{\\zeta(s)}$$\nwhich by Perron's formula suggests an asymptotic of the form\n$$ \\sum_{n \\leq x} \\Lambda_2(n) = 2 x \\log x + \\sum_\\rho \\frac{\\zeta''(\\rho)}{\\zeta'(\\rho)} \\frac{x^\\rho}{\\rho} + \\dots$$\nNow, even if there are zeroes on $\\mathrm{Re}(s)=1$, their contribution should still be smaller than the main term, and one may now be led to predict an asymptotic of the form\n$$ \\sum_{n \\leq x} \\Lambda_2(n) = 2 x \\log x + O(x) \\quad (4)$$\nand furthermore this asymptotic should be easier to prove than the prime number theorem, as it is not reliant on preventing zeroes at the edge of the critical strip.</p>\n\n<p>The functions $\\Lambda_2$ and $\\Lambda$ are related via the identity\n$$ \\frac{\\zeta''(s)}{\\zeta(s)} = -\\frac{d}{ds} (-\\frac{\\zeta'(s)}{\\zeta(s)}) +\n(-\\frac{\\zeta'(s)}{\\zeta(s)})^2$$\nwhich after inverting the Dirichlet series leads to\n$$ \\Lambda_2(n) = \\Lambda(n) \\log n + \\sum_{d|n} \\Lambda(d) \\Lambda(\\frac{n}{d}) \\quad (5)$$\nand from this and (4) we soon obtain (2) and hence (1).</p>\n\n<p>One can also think about this from a \"music of the primes\" standpoint. The explicit formula is morally resolving $\\Lambda$ into \"harmonics\" as\n$$ \\Lambda(n) \\approx 1 - \\sum_\\rho n^{\\rho-1} + \\dots \\quad (6)$$\nwhere I am deliberately being vague as to how to interpret the symbols $\\approx$, $\\sum$, and $\\dots$. The Dirichlet convolution of $1$ with itself resembles $\\log n$, while the Dirichlet convolution of $-n^{\\rho-1}$ with itself resembles $n^{\\rho-1} \\log n$; also, cross-terms such as the convolution of $1$ with $-n^{\\rho-1}$ experience significant cancellation and should be lower order. Thus the right-hand side of (5) amplifies the \"1\" harmonic of $\\Lambda$ by a factor of about $2 \\log n$, while damping out the contributions of each of the zeroes $\\rho$ due to cancellation between the two terms, leading one to (1) or (2) as a smoothed out version of the prime number theorem. Conversely, (2) combined with (5) can be viewed as an elementary way to encode (the largest terms of) the explicit formula (4), without explicit mention of zeroes.</p>\n\n<p>The most difficult part of the Erdos-Selberg elementary proof of the PNT is the passage from (1) to the prime number theorem, as here we must actually eliminate the scenario in which there is a zero on $\\mathrm{Re}(s)=1$. The key to doing so is to exploit the non-negativity of $\\Lambda$; morally, the expansion (6) prevents $\\rho$ (and hence also $\\overline{\\rho}$) from lying on $\\mathrm{Re}(s)=1$ as this would make the right-hand side RHS negative \"on average\" for certain sets of $n$. There are several ways to make this intuition precise; the standard complex-analytic proofs often proceed by testing $\\Lambda$ against $n^{\\pm it}$ and $n^{\\pm 2it}$, where $1+it$ is a hypothetical zero of $\\zeta$. Very roughly speaking, the idea of the elementary proof is to exploit not only the non-negativity of $\\Lambda$, but also upper bounds coming from sieve-theoretic bounds such as the <a href=\"https://en.wikipedia.org/wiki/Brun%E2%80%93Titchmarsh_theorem\" rel=\"noreferrer\">Brun-Titchmarsh inequality</a>, which roughly speaking tell us that $\\Lambda$ behaves \"on average\" as if it lies between $0$ and $2$. (Actually, the Selberg formula (2) already gives an adequate substitute for this inequality for the purposes of this argument.) On the other hand, from (6) we see that if there was a zero $\\rho = 1+it$ on the line $\\mathrm{Re}(s)=1$, then $\\Lambda-1$ should have an extremely large correlation with $\\mathrm{Re} n^{\\rho-1} = \\cos(t \\log n)$, and these two facts can be placed in contradiction with each other, basically because the average value of $|\\cos(t \\log n)|$ is strictly less than one. However, because the zeroes $\\rho$ are not explicitly present in the Selberg formula (2), one has to work quite a bit in the elementary proof to make this argument rigorous. In <a href=\"https://terrytao.wordpress.com/2014/10/25/a-banach-algebra-proof-of-the-prime-number-theorem/\" rel=\"noreferrer\">this blog post of mine</a> I use some Banach algebra theory to tease out the zeroes $\\rho=1+it$ from the Selberg formula (they can be interpreted as the \"Gelfand spectrum\" of a certain Banach algebra norm associated to that formula) to make this connection more explicit.</p>\n" }, { "answer_id": 259744, "score": 19, "text": "<p>Complementing Terry's nice response, let me try to explain from a more elementary point of view why Selberg's formula (1) is natural, and why it is true.</p>\n\n<p>It is natural to formulate the PNT in terms of the von Mangoldt function $\\Lambda(n)$, because it is supported on the prime powers, where it is given by the simple analytic formula $\\Lambda(p^r)=\\log p$. In addition, $\\Lambda(n)$ is intimately related to the Möbius function $\\mu(n)$ and the Riemann-zeta function via the equivalent formulae\n$$ \\Lambda=\\mu\\ast\\log,\\qquad 1\\ast\\Lambda=\\log,\\qquad \\sum_{n=1}^\\infty\\frac{\\Lambda(n)}{n^s}=-\\frac{\\zeta'}{\\zeta}(s).$$\nSelberg's insight was that a natural and slightly more accessible quantity arises if we convolve $\\mu$ not with $\\log$, but a power of $\\log$:\n$$ \\Lambda_k=\\mu\\ast\\log^k,\\qquad 1\\ast\\Lambda_k=\\log^k,\\qquad \\sum_{n=1}^\\infty\\frac{\\Lambda_k(n)}{n^s}=(-1)^k\\frac{\\zeta^{(k)}}{\\zeta}(s).$$\nIt is a nice exercise that $\\Lambda_k(n)$ is nonnegative, hence at most $\\log^k(n)$, and it is supported on numbers with at most $k$ distinct prime factors. Now (1) is essentially the same as\n$$ \\sum_{n\\leq x}\\Lambda_2(n)=2x\\log x+O(x), $$\ni.e. an analogue of the PNT for numbers with at most two distinct prime factors (which makes the sum smoother). This explains why (1) is so natural and useful to look at. It is another matter, and the ingenuity of Selberg and Erdős, that one can derive information from the distribution of numbers with at most two distinct prime factors for the distribution of prime powers.</p>\n\n<p>In general, we have the following generalization of Selberg's formula (1): \n$$ \\psi_k(x):=\\sum_{n\\leq x}\\Lambda_k(n)=x P_k(\\log x)+O(x),\\qquad k\\geq 2,$$\nwhere $P_k(t)\\in\\mathbb{R}[t]$ is a polynomial of degree $k-1$ with leading coefficient $k$. Let me sketch a proof of this more general formula. First, using the definitions,\n$$ \\psi_k(x)=\\sum_{n\\leq x}\\mu(n)L_k(x/n),$$\nwhere $L_k$ is the summatory function of $\\log^k$:\n$$ L_k(x):=\\sum_{n\\leq x}\\log^k(n).$$\nA key point is that $L_k(x)$ is a smooth sum, in fact by simple calculus \n$$ L_k(x)=xR_k(\\log x)+O(\\log^k(x)), $$\nwhere $R_k(t)\\in\\mathbb{R}[t]$ is a polynomial of degree $k$ with leading coefficient $1$. Now the idea is to compare $L_k$ with the summatory functions of the generalized divisor functions:\n$$ T_j(x):=\\sum_{n\\leq x}\\tau_j(n),\\qquad \\tau_j:=1\\ast\\dots\\ast 1. $$\nThese are easier to analyze than $\\psi_k(x)$, in fact one can prove in an elementary way that\n$$ T_j(x)=xS_j(\\log x)+O(x^{1-1/j}), \\qquad j\\geq 1,$$\nwhere $S_j(t)\\in\\mathbb{R}[t]$ is a polynomial of degree $j-1$ with leading coefficient $1/(j-1)!$. Here we used the convention that $\\tau_1:=1$, so that $T_1(x)=\\lfloor x\\rfloor=x+O(1)$ as claimed. It follows that $L_k(x)$ can be well-approximated with a linear combination of the $T_j(x)$'s:\n$$ L_k(x)=k!T_{k+1}(x)+a_kT_k(x)+\\dots+a_1T_1(x)+O(x^{1-1/k}).$$\nPlugging this back into our formula for $\\psi_k$, we get\n$$ \\psi_k(x)=\\sum_{n\\leq x}\\mu(n)\\Bigl\\{k!T_{k+1}(x/n)+a_kT_k(x/n)+\\dots+a_1T_1(x/n)\\Bigr\\}+O(x).$$\nHowever, for each $j\\geq 2$,\n$$ \\sum_{n\\leq x}\\mu(n)T_j(x/n)=\\sum_{n\\leq x}(\\mu\\ast\\tau_j)(n)=\\sum_{n\\leq x}\\tau_{j-1}(n)=T_{j-1}(x), $$\nand also\n$$ \\sum_{n\\leq x}\\mu(n)T_1(x/n)=\\sum_{n\\leq x}(\\mu\\ast 1)(n)=1, $$\nhence in fact\n$$\\begin{align*} \\psi_k(x)&amp;=k!T_k(x)+a_kT_{k-1}(x)+\\dots +a_2 T_1(x)+a_1+O(x)\\\\ \n&amp;=x\\Bigl\\{k!S_k(\\log x)+a_kS_{k-1}(\\log x)+\\dots + a_2 S_1(\\log x)\\Bigr\\} + O(x).\\end{align*} $$\nThis proves the claimed approximation for $\\psi_k(x)$ with the polynomial\n$$P_k(t):=k!S_k(t)+a_kS_{k-1}(t)+\\dots+a_2 S_1(t).$$\nThis polynomial has real coefficients, degree $k-1$ and leading coefficient $k!/(k-1)!=k$, where we used that $k\\geq 2$ throughout. The proof is complete.</p>\n" }, { "answer_id": 259792, "score": 4, "text": "<p>Two very nice answers have already been given, but I would like to add that Michel Balazard has a book in preparation on this topic (written for undergrads), in which he tries to give a deduction of the PNT from Selberg's identity which as simple as possible. Also, a student here has a nice <a href=\"https://www.eleves.ens.fr/home/guignard/pdf/pnt.pdf\" rel=\"nofollow noreferrer\">write-up</a> of the proof. It is very short, it is elementary, but it is in french.</p>\n" }, { "answer_id": 259719, "score": 52, "text": "<p>The complex-analytic proof of the prime number theorem can help inform the elementary one.</p>\n\n<p>The von Mangoldt function $\\Lambda$ is related to the Riemann zeta function $\\zeta$ by the formula\n$$ \\sum_n \\frac{\\Lambda(n)}{n^s} = -\\frac{\\zeta'(s)}{\\zeta(s)},$$\nat least in the region $\\mathrm{Re}(s) &gt; 1$. The right-hand side has a pole of residue 1 at the simple pole $s=1$ of $\\zeta$, and a pole of residue -1 at each zero $s=\\rho$ of $\\zeta$. Applying Perron's formula carefully, one eventually arrives at the von Mangoldt <a href=\"https://en.wikipedia.org/wiki/Explicit_formulae_(L-function)\" rel=\"noreferrer\">explicit formula</a></p>\n\n<p>$$ \\sum_{n \\leq x} \\Lambda(n) = x - \\sum_\\rho \\frac{x^\\rho}{\\rho} + \\dots$$\nwhere the remaining terms $\\dots$ (as well as the way in which the infinite sum over zeroes is interpreted) will be ignored for this discussion. This formula strongly suggests that zeroes $\\rho$ with real part strictly less than one will eventually only make a lower order contribution to $\\sum_{n \\leq x} \\Lambda(n)$, whereas zeroes $\\rho$ with real part equal to one would likely destroy the prime number theorem. In fact one can pursue this line of reasoning more carefully and show that the prime number theorem is equivalent to the absence of zeroes of zeta on the line $\\mathrm{Re}(s)=1$.</p>\n\n<p>But what if there was a way to make the contribution of the zeroes of zeta smaller than those of the pole, even when said zeroes lie on the line $\\mathrm{Re}(s)=1$? One way to do this is to replace the log-derivative $-\\frac{\\zeta'(s)}{\\zeta(s)}$ by the variant $\\frac{\\zeta''(s)}{\\zeta(s)}$. Assuming for simplicity that all zeroes are simple, this function still has a simple pole at every zero $\\rho$ (with residue $\\frac{\\zeta''(\\rho)}{\\zeta'(\\rho)}$), but now has a <em>double</em> pole at $s=1$ (with residue $2$). As it turns out, this variant also has a Dirichlet series representation\n$$ \\sum_n \\frac{ \\Lambda_2(n) }{n^s} = \\frac{\\zeta''(s)}{\\zeta(s)}$$\nwhich by Perron's formula suggests an asymptotic of the form\n$$ \\sum_{n \\leq x} \\Lambda_2(n) = 2 x \\log x + \\sum_\\rho \\frac{\\zeta''(\\rho)}{\\zeta'(\\rho)} \\frac{x^\\rho}{\\rho} + \\dots$$\nNow, even if there are zeroes on $\\mathrm{Re}(s)=1$, their contribution should still be smaller than the main term, and one may now be led to predict an asymptotic of the form\n$$ \\sum_{n \\leq x} \\Lambda_2(n) = 2 x \\log x + O(x) \\quad (4)$$\nand furthermore this asymptotic should be easier to prove than the prime number theorem, as it is not reliant on preventing zeroes at the edge of the critical strip.</p>\n\n<p>The functions $\\Lambda_2$ and $\\Lambda$ are related via the identity\n$$ \\frac{\\zeta''(s)}{\\zeta(s)} = -\\frac{d}{ds} (-\\frac{\\zeta'(s)}{\\zeta(s)}) +\n(-\\frac{\\zeta'(s)}{\\zeta(s)})^2$$\nwhich after inverting the Dirichlet series leads to\n$$ \\Lambda_2(n) = \\Lambda(n) \\log n + \\sum_{d|n} \\Lambda(d) \\Lambda(\\frac{n}{d}) \\quad (5)$$\nand from this and (4) we soon obtain (2) and hence (1).</p>\n\n<p>One can also think about this from a \"music of the primes\" standpoint. The explicit formula is morally resolving $\\Lambda$ into \"harmonics\" as\n$$ \\Lambda(n) \\approx 1 - \\sum_\\rho n^{\\rho-1} + \\dots \\quad (6)$$\nwhere I am deliberately being vague as to how to interpret the symbols $\\approx$, $\\sum$, and $\\dots$. The Dirichlet convolution of $1$ with itself resembles $\\log n$, while the Dirichlet convolution of $-n^{\\rho-1}$ with itself resembles $n^{\\rho-1} \\log n$; also, cross-terms such as the convolution of $1$ with $-n^{\\rho-1}$ experience significant cancellation and should be lower order. Thus the right-hand side of (5) amplifies the \"1\" harmonic of $\\Lambda$ by a factor of about $2 \\log n$, while damping out the contributions of each of the zeroes $\\rho$ due to cancellation between the two terms, leading one to (1) or (2) as a smoothed out version of the prime number theorem. Conversely, (2) combined with (5) can be viewed as an elementary way to encode (the largest terms of) the explicit formula (4), without explicit mention of zeroes.</p>\n\n<p>The most difficult part of the Erdos-Selberg elementary proof of the PNT is the passage from (1) to the prime number theorem, as here we must actually eliminate the scenario in which there is a zero on $\\mathrm{Re}(s)=1$. The key to doing so is to exploit the non-negativity of $\\Lambda$; morally, the expansion (6) prevents $\\rho$ (and hence also $\\overline{\\rho}$) from lying on $\\mathrm{Re}(s)=1$ as this would make the right-hand side RHS negative \"on average\" for certain sets of $n$. There are several ways to make this intuition precise; the standard complex-analytic proofs often proceed by testing $\\Lambda$ against $n^{\\pm it}$ and $n^{\\pm 2it}$, where $1+it$ is a hypothetical zero of $\\zeta$. Very roughly speaking, the idea of the elementary proof is to exploit not only the non-negativity of $\\Lambda$, but also upper bounds coming from sieve-theoretic bounds such as the <a href=\"https://en.wikipedia.org/wiki/Brun%E2%80%93Titchmarsh_theorem\" rel=\"noreferrer\">Brun-Titchmarsh inequality</a>, which roughly speaking tell us that $\\Lambda$ behaves \"on average\" as if it lies between $0$ and $2$. (Actually, the Selberg formula (2) already gives an adequate substitute for this inequality for the purposes of this argument.) On the other hand, from (6) we see that if there was a zero $\\rho = 1+it$ on the line $\\mathrm{Re}(s)=1$, then $\\Lambda-1$ should have an extremely large correlation with $\\mathrm{Re} n^{\\rho-1} = \\cos(t \\log n)$, and these two facts can be placed in contradiction with each other, basically because the average value of $|\\cos(t \\log n)|$ is strictly less than one. However, because the zeroes $\\rho$ are not explicitly present in the Selberg formula (2), one has to work quite a bit in the elementary proof to make this argument rigorous. In <a href=\"https://terrytao.wordpress.com/2014/10/25/a-banach-algebra-proof-of-the-prime-number-theorem/\" rel=\"noreferrer\">this blog post of mine</a> I use some Banach algebra theory to tease out the zeroes $\\rho=1+it$ from the Selberg formula (they can be interpreted as the \"Gelfand spectrum\" of a certain Banach algebra norm associated to that formula) to make this connection more explicit.</p>\n" }, { "answer_id": 259744, "score": 19, "text": "<p>Complementing Terry's nice response, let me try to explain from a more elementary point of view why Selberg's formula (1) is natural, and why it is true.</p>\n\n<p>It is natural to formulate the PNT in terms of the von Mangoldt function $\\Lambda(n)$, because it is supported on the prime powers, where it is given by the simple analytic formula $\\Lambda(p^r)=\\log p$. In addition, $\\Lambda(n)$ is intimately related to the Möbius function $\\mu(n)$ and the Riemann-zeta function via the equivalent formulae\n$$ \\Lambda=\\mu\\ast\\log,\\qquad 1\\ast\\Lambda=\\log,\\qquad \\sum_{n=1}^\\infty\\frac{\\Lambda(n)}{n^s}=-\\frac{\\zeta'}{\\zeta}(s).$$\nSelberg's insight was that a natural and slightly more accessible quantity arises if we convolve $\\mu$ not with $\\log$, but a power of $\\log$:\n$$ \\Lambda_k=\\mu\\ast\\log^k,\\qquad 1\\ast\\Lambda_k=\\log^k,\\qquad \\sum_{n=1}^\\infty\\frac{\\Lambda_k(n)}{n^s}=(-1)^k\\frac{\\zeta^{(k)}}{\\zeta}(s).$$\nIt is a nice exercise that $\\Lambda_k(n)$ is nonnegative, hence at most $\\log^k(n)$, and it is supported on numbers with at most $k$ distinct prime factors. Now (1) is essentially the same as\n$$ \\sum_{n\\leq x}\\Lambda_2(n)=2x\\log x+O(x), $$\ni.e. an analogue of the PNT for numbers with at most two distinct prime factors (which makes the sum smoother). This explains why (1) is so natural and useful to look at. It is another matter, and the ingenuity of Selberg and Erdős, that one can derive information from the distribution of numbers with at most two distinct prime factors for the distribution of prime powers.</p>\n\n<p>In general, we have the following generalization of Selberg's formula (1): \n$$ \\psi_k(x):=\\sum_{n\\leq x}\\Lambda_k(n)=x P_k(\\log x)+O(x),\\qquad k\\geq 2,$$\nwhere $P_k(t)\\in\\mathbb{R}[t]$ is a polynomial of degree $k-1$ with leading coefficient $k$. Let me sketch a proof of this more general formula. First, using the definitions,\n$$ \\psi_k(x)=\\sum_{n\\leq x}\\mu(n)L_k(x/n),$$\nwhere $L_k$ is the summatory function of $\\log^k$:\n$$ L_k(x):=\\sum_{n\\leq x}\\log^k(n).$$\nA key point is that $L_k(x)$ is a smooth sum, in fact by simple calculus \n$$ L_k(x)=xR_k(\\log x)+O(\\log^k(x)), $$\nwhere $R_k(t)\\in\\mathbb{R}[t]$ is a polynomial of degree $k$ with leading coefficient $1$. Now the idea is to compare $L_k$ with the summatory functions of the generalized divisor functions:\n$$ T_j(x):=\\sum_{n\\leq x}\\tau_j(n),\\qquad \\tau_j:=1\\ast\\dots\\ast 1. $$\nThese are easier to analyze than $\\psi_k(x)$, in fact one can prove in an elementary way that\n$$ T_j(x)=xS_j(\\log x)+O(x^{1-1/j}), \\qquad j\\geq 1,$$\nwhere $S_j(t)\\in\\mathbb{R}[t]$ is a polynomial of degree $j-1$ with leading coefficient $1/(j-1)!$. Here we used the convention that $\\tau_1:=1$, so that $T_1(x)=\\lfloor x\\rfloor=x+O(1)$ as claimed. It follows that $L_k(x)$ can be well-approximated with a linear combination of the $T_j(x)$'s:\n$$ L_k(x)=k!T_{k+1}(x)+a_kT_k(x)+\\dots+a_1T_1(x)+O(x^{1-1/k}).$$\nPlugging this back into our formula for $\\psi_k$, we get\n$$ \\psi_k(x)=\\sum_{n\\leq x}\\mu(n)\\Bigl\\{k!T_{k+1}(x/n)+a_kT_k(x/n)+\\dots+a_1T_1(x/n)\\Bigr\\}+O(x).$$\nHowever, for each $j\\geq 2$,\n$$ \\sum_{n\\leq x}\\mu(n)T_j(x/n)=\\sum_{n\\leq x}(\\mu\\ast\\tau_j)(n)=\\sum_{n\\leq x}\\tau_{j-1}(n)=T_{j-1}(x), $$\nand also\n$$ \\sum_{n\\leq x}\\mu(n)T_1(x/n)=\\sum_{n\\leq x}(\\mu\\ast 1)(n)=1, $$\nhence in fact\n$$\\begin{align*} \\psi_k(x)&amp;=k!T_k(x)+a_kT_{k-1}(x)+\\dots +a_2 T_1(x)+a_1+O(x)\\\\ \n&amp;=x\\Bigl\\{k!S_k(\\log x)+a_kS_{k-1}(\\log x)+\\dots + a_2 S_1(\\log x)\\Bigr\\} + O(x).\\end{align*} $$\nThis proves the claimed approximation for $\\psi_k(x)$ with the polynomial\n$$P_k(t):=k!S_k(t)+a_kS_{k-1}(t)+\\dots+a_2 S_1(t).$$\nThis polynomial has real coefficients, degree $k-1$ and leading coefficient $k!/(k-1)!=k$, where we used that $k\\geq 2$ throughout. The proof is complete.</p>\n" }, { "answer_id": 259792, "score": 4, "text": "<p>Two very nice answers have already been given, but I would like to add that Michel Balazard has a book in preparation on this topic (written for undergrads), in which he tries to give a deduction of the PNT from Selberg's identity which as simple as possible. Also, a student here has a nice <a href=\"https://www.eleves.ens.fr/home/guignard/pdf/pnt.pdf\" rel=\"nofollow noreferrer\">write-up</a> of the proof. It is very short, it is elementary, but it is in french.</p>\n" }, { "answer_id": 259719, "score": 52, "text": "<p>The complex-analytic proof of the prime number theorem can help inform the elementary one.</p>\n\n<p>The von Mangoldt function $\\Lambda$ is related to the Riemann zeta function $\\zeta$ by the formula\n$$ \\sum_n \\frac{\\Lambda(n)}{n^s} = -\\frac{\\zeta'(s)}{\\zeta(s)},$$\nat least in the region $\\mathrm{Re}(s) &gt; 1$. The right-hand side has a pole of residue 1 at the simple pole $s=1$ of $\\zeta$, and a pole of residue -1 at each zero $s=\\rho$ of $\\zeta$. Applying Perron's formula carefully, one eventually arrives at the von Mangoldt <a href=\"https://en.wikipedia.org/wiki/Explicit_formulae_(L-function)\" rel=\"noreferrer\">explicit formula</a></p>\n\n<p>$$ \\sum_{n \\leq x} \\Lambda(n) = x - \\sum_\\rho \\frac{x^\\rho}{\\rho} + \\dots$$\nwhere the remaining terms $\\dots$ (as well as the way in which the infinite sum over zeroes is interpreted) will be ignored for this discussion. This formula strongly suggests that zeroes $\\rho$ with real part strictly less than one will eventually only make a lower order contribution to $\\sum_{n \\leq x} \\Lambda(n)$, whereas zeroes $\\rho$ with real part equal to one would likely destroy the prime number theorem. In fact one can pursue this line of reasoning more carefully and show that the prime number theorem is equivalent to the absence of zeroes of zeta on the line $\\mathrm{Re}(s)=1$.</p>\n\n<p>But what if there was a way to make the contribution of the zeroes of zeta smaller than those of the pole, even when said zeroes lie on the line $\\mathrm{Re}(s)=1$? One way to do this is to replace the log-derivative $-\\frac{\\zeta'(s)}{\\zeta(s)}$ by the variant $\\frac{\\zeta''(s)}{\\zeta(s)}$. Assuming for simplicity that all zeroes are simple, this function still has a simple pole at every zero $\\rho$ (with residue $\\frac{\\zeta''(\\rho)}{\\zeta'(\\rho)}$), but now has a <em>double</em> pole at $s=1$ (with residue $2$). As it turns out, this variant also has a Dirichlet series representation\n$$ \\sum_n \\frac{ \\Lambda_2(n) }{n^s} = \\frac{\\zeta''(s)}{\\zeta(s)}$$\nwhich by Perron's formula suggests an asymptotic of the form\n$$ \\sum_{n \\leq x} \\Lambda_2(n) = 2 x \\log x + \\sum_\\rho \\frac{\\zeta''(\\rho)}{\\zeta'(\\rho)} \\frac{x^\\rho}{\\rho} + \\dots$$\nNow, even if there are zeroes on $\\mathrm{Re}(s)=1$, their contribution should still be smaller than the main term, and one may now be led to predict an asymptotic of the form\n$$ \\sum_{n \\leq x} \\Lambda_2(n) = 2 x \\log x + O(x) \\quad (4)$$\nand furthermore this asymptotic should be easier to prove than the prime number theorem, as it is not reliant on preventing zeroes at the edge of the critical strip.</p>\n\n<p>The functions $\\Lambda_2$ and $\\Lambda$ are related via the identity\n$$ \\frac{\\zeta''(s)}{\\zeta(s)} = -\\frac{d}{ds} (-\\frac{\\zeta'(s)}{\\zeta(s)}) +\n(-\\frac{\\zeta'(s)}{\\zeta(s)})^2$$\nwhich after inverting the Dirichlet series leads to\n$$ \\Lambda_2(n) = \\Lambda(n) \\log n + \\sum_{d|n} \\Lambda(d) \\Lambda(\\frac{n}{d}) \\quad (5)$$\nand from this and (4) we soon obtain (2) and hence (1).</p>\n\n<p>One can also think about this from a \"music of the primes\" standpoint. The explicit formula is morally resolving $\\Lambda$ into \"harmonics\" as\n$$ \\Lambda(n) \\approx 1 - \\sum_\\rho n^{\\rho-1} + \\dots \\quad (6)$$\nwhere I am deliberately being vague as to how to interpret the symbols $\\approx$, $\\sum$, and $\\dots$. The Dirichlet convolution of $1$ with itself resembles $\\log n$, while the Dirichlet convolution of $-n^{\\rho-1}$ with itself resembles $n^{\\rho-1} \\log n$; also, cross-terms such as the convolution of $1$ with $-n^{\\rho-1}$ experience significant cancellation and should be lower order. Thus the right-hand side of (5) amplifies the \"1\" harmonic of $\\Lambda$ by a factor of about $2 \\log n$, while damping out the contributions of each of the zeroes $\\rho$ due to cancellation between the two terms, leading one to (1) or (2) as a smoothed out version of the prime number theorem. Conversely, (2) combined with (5) can be viewed as an elementary way to encode (the largest terms of) the explicit formula (4), without explicit mention of zeroes.</p>\n\n<p>The most difficult part of the Erdos-Selberg elementary proof of the PNT is the passage from (1) to the prime number theorem, as here we must actually eliminate the scenario in which there is a zero on $\\mathrm{Re}(s)=1$. The key to doing so is to exploit the non-negativity of $\\Lambda$; morally, the expansion (6) prevents $\\rho$ (and hence also $\\overline{\\rho}$) from lying on $\\mathrm{Re}(s)=1$ as this would make the right-hand side RHS negative \"on average\" for certain sets of $n$. There are several ways to make this intuition precise; the standard complex-analytic proofs often proceed by testing $\\Lambda$ against $n^{\\pm it}$ and $n^{\\pm 2it}$, where $1+it$ is a hypothetical zero of $\\zeta$. Very roughly speaking, the idea of the elementary proof is to exploit not only the non-negativity of $\\Lambda$, but also upper bounds coming from sieve-theoretic bounds such as the <a href=\"https://en.wikipedia.org/wiki/Brun%E2%80%93Titchmarsh_theorem\" rel=\"noreferrer\">Brun-Titchmarsh inequality</a>, which roughly speaking tell us that $\\Lambda$ behaves \"on average\" as if it lies between $0$ and $2$. (Actually, the Selberg formula (2) already gives an adequate substitute for this inequality for the purposes of this argument.) On the other hand, from (6) we see that if there was a zero $\\rho = 1+it$ on the line $\\mathrm{Re}(s)=1$, then $\\Lambda-1$ should have an extremely large correlation with $\\mathrm{Re} n^{\\rho-1} = \\cos(t \\log n)$, and these two facts can be placed in contradiction with each other, basically because the average value of $|\\cos(t \\log n)|$ is strictly less than one. However, because the zeroes $\\rho$ are not explicitly present in the Selberg formula (2), one has to work quite a bit in the elementary proof to make this argument rigorous. In <a href=\"https://terrytao.wordpress.com/2014/10/25/a-banach-algebra-proof-of-the-prime-number-theorem/\" rel=\"noreferrer\">this blog post of mine</a> I use some Banach algebra theory to tease out the zeroes $\\rho=1+it$ from the Selberg formula (they can be interpreted as the \"Gelfand spectrum\" of a certain Banach algebra norm associated to that formula) to make this connection more explicit.</p>\n" }, { "answer_id": 259744, "score": 19, "text": "<p>Complementing Terry's nice response, let me try to explain from a more elementary point of view why Selberg's formula (1) is natural, and why it is true.</p>\n\n<p>It is natural to formulate the PNT in terms of the von Mangoldt function $\\Lambda(n)$, because it is supported on the prime powers, where it is given by the simple analytic formula $\\Lambda(p^r)=\\log p$. In addition, $\\Lambda(n)$ is intimately related to the Möbius function $\\mu(n)$ and the Riemann-zeta function via the equivalent formulae\n$$ \\Lambda=\\mu\\ast\\log,\\qquad 1\\ast\\Lambda=\\log,\\qquad \\sum_{n=1}^\\infty\\frac{\\Lambda(n)}{n^s}=-\\frac{\\zeta'}{\\zeta}(s).$$\nSelberg's insight was that a natural and slightly more accessible quantity arises if we convolve $\\mu$ not with $\\log$, but a power of $\\log$:\n$$ \\Lambda_k=\\mu\\ast\\log^k,\\qquad 1\\ast\\Lambda_k=\\log^k,\\qquad \\sum_{n=1}^\\infty\\frac{\\Lambda_k(n)}{n^s}=(-1)^k\\frac{\\zeta^{(k)}}{\\zeta}(s).$$\nIt is a nice exercise that $\\Lambda_k(n)$ is nonnegative, hence at most $\\log^k(n)$, and it is supported on numbers with at most $k$ distinct prime factors. Now (1) is essentially the same as\n$$ \\sum_{n\\leq x}\\Lambda_2(n)=2x\\log x+O(x), $$\ni.e. an analogue of the PNT for numbers with at most two distinct prime factors (which makes the sum smoother). This explains why (1) is so natural and useful to look at. It is another matter, and the ingenuity of Selberg and Erdős, that one can derive information from the distribution of numbers with at most two distinct prime factors for the distribution of prime powers.</p>\n\n<p>In general, we have the following generalization of Selberg's formula (1): \n$$ \\psi_k(x):=\\sum_{n\\leq x}\\Lambda_k(n)=x P_k(\\log x)+O(x),\\qquad k\\geq 2,$$\nwhere $P_k(t)\\in\\mathbb{R}[t]$ is a polynomial of degree $k-1$ with leading coefficient $k$. Let me sketch a proof of this more general formula. First, using the definitions,\n$$ \\psi_k(x)=\\sum_{n\\leq x}\\mu(n)L_k(x/n),$$\nwhere $L_k$ is the summatory function of $\\log^k$:\n$$ L_k(x):=\\sum_{n\\leq x}\\log^k(n).$$\nA key point is that $L_k(x)$ is a smooth sum, in fact by simple calculus \n$$ L_k(x)=xR_k(\\log x)+O(\\log^k(x)), $$\nwhere $R_k(t)\\in\\mathbb{R}[t]$ is a polynomial of degree $k$ with leading coefficient $1$. Now the idea is to compare $L_k$ with the summatory functions of the generalized divisor functions:\n$$ T_j(x):=\\sum_{n\\leq x}\\tau_j(n),\\qquad \\tau_j:=1\\ast\\dots\\ast 1. $$\nThese are easier to analyze than $\\psi_k(x)$, in fact one can prove in an elementary way that\n$$ T_j(x)=xS_j(\\log x)+O(x^{1-1/j}), \\qquad j\\geq 1,$$\nwhere $S_j(t)\\in\\mathbb{R}[t]$ is a polynomial of degree $j-1$ with leading coefficient $1/(j-1)!$. Here we used the convention that $\\tau_1:=1$, so that $T_1(x)=\\lfloor x\\rfloor=x+O(1)$ as claimed. It follows that $L_k(x)$ can be well-approximated with a linear combination of the $T_j(x)$'s:\n$$ L_k(x)=k!T_{k+1}(x)+a_kT_k(x)+\\dots+a_1T_1(x)+O(x^{1-1/k}).$$\nPlugging this back into our formula for $\\psi_k$, we get\n$$ \\psi_k(x)=\\sum_{n\\leq x}\\mu(n)\\Bigl\\{k!T_{k+1}(x/n)+a_kT_k(x/n)+\\dots+a_1T_1(x/n)\\Bigr\\}+O(x).$$\nHowever, for each $j\\geq 2$,\n$$ \\sum_{n\\leq x}\\mu(n)T_j(x/n)=\\sum_{n\\leq x}(\\mu\\ast\\tau_j)(n)=\\sum_{n\\leq x}\\tau_{j-1}(n)=T_{j-1}(x), $$\nand also\n$$ \\sum_{n\\leq x}\\mu(n)T_1(x/n)=\\sum_{n\\leq x}(\\mu\\ast 1)(n)=1, $$\nhence in fact\n$$\\begin{align*} \\psi_k(x)&amp;=k!T_k(x)+a_kT_{k-1}(x)+\\dots +a_2 T_1(x)+a_1+O(x)\\\\ \n&amp;=x\\Bigl\\{k!S_k(\\log x)+a_kS_{k-1}(\\log x)+\\dots + a_2 S_1(\\log x)\\Bigr\\} + O(x).\\end{align*} $$\nThis proves the claimed approximation for $\\psi_k(x)$ with the polynomial\n$$P_k(t):=k!S_k(t)+a_kS_{k-1}(t)+\\dots+a_2 S_1(t).$$\nThis polynomial has real coefficients, degree $k-1$ and leading coefficient $k!/(k-1)!=k$, where we used that $k\\geq 2$ throughout. The proof is complete.</p>\n" }, { "answer_id": 259792, "score": 4, "text": "<p>Two very nice answers have already been given, but I would like to add that Michel Balazard has a book in preparation on this topic (written for undergrads), in which he tries to give a deduction of the PNT from Selberg's identity which as simple as possible. Also, a student here has a nice <a href=\"https://www.eleves.ens.fr/home/guignard/pdf/pnt.pdf\" rel=\"nofollow noreferrer\">write-up</a> of the proof. It is very short, it is elementary, but it is in french.</p>\n" }, { "answer_id": 259719, "score": 52, "text": "<p>The complex-analytic proof of the prime number theorem can help inform the elementary one.</p>\n\n<p>The von Mangoldt function $\\Lambda$ is related to the Riemann zeta function $\\zeta$ by the formula\n$$ \\sum_n \\frac{\\Lambda(n)}{n^s} = -\\frac{\\zeta'(s)}{\\zeta(s)},$$\nat least in the region $\\mathrm{Re}(s) &gt; 1$. The right-hand side has a pole of residue 1 at the simple pole $s=1$ of $\\zeta$, and a pole of residue -1 at each zero $s=\\rho$ of $\\zeta$. Applying Perron's formula carefully, one eventually arrives at the von Mangoldt <a href=\"https://en.wikipedia.org/wiki/Explicit_formulae_(L-function)\" rel=\"noreferrer\">explicit formula</a></p>\n\n<p>$$ \\sum_{n \\leq x} \\Lambda(n) = x - \\sum_\\rho \\frac{x^\\rho}{\\rho} + \\dots$$\nwhere the remaining terms $\\dots$ (as well as the way in which the infinite sum over zeroes is interpreted) will be ignored for this discussion. This formula strongly suggests that zeroes $\\rho$ with real part strictly less than one will eventually only make a lower order contribution to $\\sum_{n \\leq x} \\Lambda(n)$, whereas zeroes $\\rho$ with real part equal to one would likely destroy the prime number theorem. In fact one can pursue this line of reasoning more carefully and show that the prime number theorem is equivalent to the absence of zeroes of zeta on the line $\\mathrm{Re}(s)=1$.</p>\n\n<p>But what if there was a way to make the contribution of the zeroes of zeta smaller than those of the pole, even when said zeroes lie on the line $\\mathrm{Re}(s)=1$? One way to do this is to replace the log-derivative $-\\frac{\\zeta'(s)}{\\zeta(s)}$ by the variant $\\frac{\\zeta''(s)}{\\zeta(s)}$. Assuming for simplicity that all zeroes are simple, this function still has a simple pole at every zero $\\rho$ (with residue $\\frac{\\zeta''(\\rho)}{\\zeta'(\\rho)}$), but now has a <em>double</em> pole at $s=1$ (with residue $2$). As it turns out, this variant also has a Dirichlet series representation\n$$ \\sum_n \\frac{ \\Lambda_2(n) }{n^s} = \\frac{\\zeta''(s)}{\\zeta(s)}$$\nwhich by Perron's formula suggests an asymptotic of the form\n$$ \\sum_{n \\leq x} \\Lambda_2(n) = 2 x \\log x + \\sum_\\rho \\frac{\\zeta''(\\rho)}{\\zeta'(\\rho)} \\frac{x^\\rho}{\\rho} + \\dots$$\nNow, even if there are zeroes on $\\mathrm{Re}(s)=1$, their contribution should still be smaller than the main term, and one may now be led to predict an asymptotic of the form\n$$ \\sum_{n \\leq x} \\Lambda_2(n) = 2 x \\log x + O(x) \\quad (4)$$\nand furthermore this asymptotic should be easier to prove than the prime number theorem, as it is not reliant on preventing zeroes at the edge of the critical strip.</p>\n\n<p>The functions $\\Lambda_2$ and $\\Lambda$ are related via the identity\n$$ \\frac{\\zeta''(s)}{\\zeta(s)} = -\\frac{d}{ds} (-\\frac{\\zeta'(s)}{\\zeta(s)}) +\n(-\\frac{\\zeta'(s)}{\\zeta(s)})^2$$\nwhich after inverting the Dirichlet series leads to\n$$ \\Lambda_2(n) = \\Lambda(n) \\log n + \\sum_{d|n} \\Lambda(d) \\Lambda(\\frac{n}{d}) \\quad (5)$$\nand from this and (4) we soon obtain (2) and hence (1).</p>\n\n<p>One can also think about this from a \"music of the primes\" standpoint. The explicit formula is morally resolving $\\Lambda$ into \"harmonics\" as\n$$ \\Lambda(n) \\approx 1 - \\sum_\\rho n^{\\rho-1} + \\dots \\quad (6)$$\nwhere I am deliberately being vague as to how to interpret the symbols $\\approx$, $\\sum$, and $\\dots$. The Dirichlet convolution of $1$ with itself resembles $\\log n$, while the Dirichlet convolution of $-n^{\\rho-1}$ with itself resembles $n^{\\rho-1} \\log n$; also, cross-terms such as the convolution of $1$ with $-n^{\\rho-1}$ experience significant cancellation and should be lower order. Thus the right-hand side of (5) amplifies the \"1\" harmonic of $\\Lambda$ by a factor of about $2 \\log n$, while damping out the contributions of each of the zeroes $\\rho$ due to cancellation between the two terms, leading one to (1) or (2) as a smoothed out version of the prime number theorem. Conversely, (2) combined with (5) can be viewed as an elementary way to encode (the largest terms of) the explicit formula (4), without explicit mention of zeroes.</p>\n\n<p>The most difficult part of the Erdos-Selberg elementary proof of the PNT is the passage from (1) to the prime number theorem, as here we must actually eliminate the scenario in which there is a zero on $\\mathrm{Re}(s)=1$. The key to doing so is to exploit the non-negativity of $\\Lambda$; morally, the expansion (6) prevents $\\rho$ (and hence also $\\overline{\\rho}$) from lying on $\\mathrm{Re}(s)=1$ as this would make the right-hand side RHS negative \"on average\" for certain sets of $n$. There are several ways to make this intuition precise; the standard complex-analytic proofs often proceed by testing $\\Lambda$ against $n^{\\pm it}$ and $n^{\\pm 2it}$, where $1+it$ is a hypothetical zero of $\\zeta$. Very roughly speaking, the idea of the elementary proof is to exploit not only the non-negativity of $\\Lambda$, but also upper bounds coming from sieve-theoretic bounds such as the <a href=\"https://en.wikipedia.org/wiki/Brun%E2%80%93Titchmarsh_theorem\" rel=\"noreferrer\">Brun-Titchmarsh inequality</a>, which roughly speaking tell us that $\\Lambda$ behaves \"on average\" as if it lies between $0$ and $2$. (Actually, the Selberg formula (2) already gives an adequate substitute for this inequality for the purposes of this argument.) On the other hand, from (6) we see that if there was a zero $\\rho = 1+it$ on the line $\\mathrm{Re}(s)=1$, then $\\Lambda-1$ should have an extremely large correlation with $\\mathrm{Re} n^{\\rho-1} = \\cos(t \\log n)$, and these two facts can be placed in contradiction with each other, basically because the average value of $|\\cos(t \\log n)|$ is strictly less than one. However, because the zeroes $\\rho$ are not explicitly present in the Selberg formula (2), one has to work quite a bit in the elementary proof to make this argument rigorous. In <a href=\"https://terrytao.wordpress.com/2014/10/25/a-banach-algebra-proof-of-the-prime-number-theorem/\" rel=\"noreferrer\">this blog post of mine</a> I use some Banach algebra theory to tease out the zeroes $\\rho=1+it$ from the Selberg formula (they can be interpreted as the \"Gelfand spectrum\" of a certain Banach algebra norm associated to that formula) to make this connection more explicit.</p>\n" }, { "answer_id": 259744, "score": 19, "text": "<p>Complementing Terry's nice response, let me try to explain from a more elementary point of view why Selberg's formula (1) is natural, and why it is true.</p>\n\n<p>It is natural to formulate the PNT in terms of the von Mangoldt function $\\Lambda(n)$, because it is supported on the prime powers, where it is given by the simple analytic formula $\\Lambda(p^r)=\\log p$. In addition, $\\Lambda(n)$ is intimately related to the Möbius function $\\mu(n)$ and the Riemann-zeta function via the equivalent formulae\n$$ \\Lambda=\\mu\\ast\\log,\\qquad 1\\ast\\Lambda=\\log,\\qquad \\sum_{n=1}^\\infty\\frac{\\Lambda(n)}{n^s}=-\\frac{\\zeta'}{\\zeta}(s).$$\nSelberg's insight was that a natural and slightly more accessible quantity arises if we convolve $\\mu$ not with $\\log$, but a power of $\\log$:\n$$ \\Lambda_k=\\mu\\ast\\log^k,\\qquad 1\\ast\\Lambda_k=\\log^k,\\qquad \\sum_{n=1}^\\infty\\frac{\\Lambda_k(n)}{n^s}=(-1)^k\\frac{\\zeta^{(k)}}{\\zeta}(s).$$\nIt is a nice exercise that $\\Lambda_k(n)$ is nonnegative, hence at most $\\log^k(n)$, and it is supported on numbers with at most $k$ distinct prime factors. Now (1) is essentially the same as\n$$ \\sum_{n\\leq x}\\Lambda_2(n)=2x\\log x+O(x), $$\ni.e. an analogue of the PNT for numbers with at most two distinct prime factors (which makes the sum smoother). This explains why (1) is so natural and useful to look at. It is another matter, and the ingenuity of Selberg and Erdős, that one can derive information from the distribution of numbers with at most two distinct prime factors for the distribution of prime powers.</p>\n\n<p>In general, we have the following generalization of Selberg's formula (1): \n$$ \\psi_k(x):=\\sum_{n\\leq x}\\Lambda_k(n)=x P_k(\\log x)+O(x),\\qquad k\\geq 2,$$\nwhere $P_k(t)\\in\\mathbb{R}[t]$ is a polynomial of degree $k-1$ with leading coefficient $k$. Let me sketch a proof of this more general formula. First, using the definitions,\n$$ \\psi_k(x)=\\sum_{n\\leq x}\\mu(n)L_k(x/n),$$\nwhere $L_k$ is the summatory function of $\\log^k$:\n$$ L_k(x):=\\sum_{n\\leq x}\\log^k(n).$$\nA key point is that $L_k(x)$ is a smooth sum, in fact by simple calculus \n$$ L_k(x)=xR_k(\\log x)+O(\\log^k(x)), $$\nwhere $R_k(t)\\in\\mathbb{R}[t]$ is a polynomial of degree $k$ with leading coefficient $1$. Now the idea is to compare $L_k$ with the summatory functions of the generalized divisor functions:\n$$ T_j(x):=\\sum_{n\\leq x}\\tau_j(n),\\qquad \\tau_j:=1\\ast\\dots\\ast 1. $$\nThese are easier to analyze than $\\psi_k(x)$, in fact one can prove in an elementary way that\n$$ T_j(x)=xS_j(\\log x)+O(x^{1-1/j}), \\qquad j\\geq 1,$$\nwhere $S_j(t)\\in\\mathbb{R}[t]$ is a polynomial of degree $j-1$ with leading coefficient $1/(j-1)!$. Here we used the convention that $\\tau_1:=1$, so that $T_1(x)=\\lfloor x\\rfloor=x+O(1)$ as claimed. It follows that $L_k(x)$ can be well-approximated with a linear combination of the $T_j(x)$'s:\n$$ L_k(x)=k!T_{k+1}(x)+a_kT_k(x)+\\dots+a_1T_1(x)+O(x^{1-1/k}).$$\nPlugging this back into our formula for $\\psi_k$, we get\n$$ \\psi_k(x)=\\sum_{n\\leq x}\\mu(n)\\Bigl\\{k!T_{k+1}(x/n)+a_kT_k(x/n)+\\dots+a_1T_1(x/n)\\Bigr\\}+O(x).$$\nHowever, for each $j\\geq 2$,\n$$ \\sum_{n\\leq x}\\mu(n)T_j(x/n)=\\sum_{n\\leq x}(\\mu\\ast\\tau_j)(n)=\\sum_{n\\leq x}\\tau_{j-1}(n)=T_{j-1}(x), $$\nand also\n$$ \\sum_{n\\leq x}\\mu(n)T_1(x/n)=\\sum_{n\\leq x}(\\mu\\ast 1)(n)=1, $$\nhence in fact\n$$\\begin{align*} \\psi_k(x)&amp;=k!T_k(x)+a_kT_{k-1}(x)+\\dots +a_2 T_1(x)+a_1+O(x)\\\\ \n&amp;=x\\Bigl\\{k!S_k(\\log x)+a_kS_{k-1}(\\log x)+\\dots + a_2 S_1(\\log x)\\Bigr\\} + O(x).\\end{align*} $$\nThis proves the claimed approximation for $\\psi_k(x)$ with the polynomial\n$$P_k(t):=k!S_k(t)+a_kS_{k-1}(t)+\\dots+a_2 S_1(t).$$\nThis polynomial has real coefficients, degree $k-1$ and leading coefficient $k!/(k-1)!=k$, where we used that $k\\geq 2$ throughout. The proof is complete.</p>\n" }, { "answer_id": 259792, "score": 4, "text": "<p>Two very nice answers have already been given, but I would like to add that Michel Balazard has a book in preparation on this topic (written for undergrads), in which he tries to give a deduction of the PNT from Selberg's identity which as simple as possible. Also, a student here has a nice <a href=\"https://www.eleves.ens.fr/home/guignard/pdf/pnt.pdf\" rel=\"nofollow noreferrer\">write-up</a> of the proof. It is very short, it is elementary, but it is in french.</p>\n" } ]
Question: Ideas in the elementary proof of the prime number theorem (Selberg / Erdős) Answer 1: <p>The complex-analytic proof of the prime number theorem can help inform the elementary one.</p> <p>The von Mangoldt function $\Lambda$ is related to the Riemann zeta function $\zeta$ by the formula $$ \sum_n \frac{\Lambda(n)}{n^s} = -\frac{\zeta'(s)}{\zeta(s)},$$ at least in the region $\mathrm{Re}(s) &gt; 1$. The right-hand side has a pole of residue 1 at the simple pole $s=1$ of $\zeta$, and a pole of residue -1 at each zero $s=\rho$ of $\zeta$. Applying Perron's formula carefully, one eventually arrives at the von Mangoldt <a href="https://en.wikipedia.org/wiki/Explicit_formulae_(L-function)" rel="noreferrer">explicit formula</a></p> <p>$$ \sum_{n \leq x} \Lambda(n) = x - \sum_\rho \frac{x^\rho}{\rho} + \dots$$ where the remaining terms $\dots$ (as well as the way in which the infinite sum over zeroes is interpreted) will be ignored for this discussion. This formula strongly suggests that zeroes $\rho$ with real part strictly less than one will eventually only make a lower order contribution to $\sum_{n \leq x} \Lambda(n)$, whereas zeroes $\rho$ with real part equal to one would likely destroy the prime number theorem. In fact one can pursue this line of reasoning more carefully and show that the prime number theorem is equivalent to the absence of zeroes of zeta on the line $\mathrm{Re}(s)=1$.</p> <p>But what if there was a way to make the contribution of the zeroes of zeta smaller than those of the pole, even when said zeroes lie on the line $\mathrm{Re}(s)=1$? One way to do this is to replace the log-derivative $-\frac{\zeta'(s)}{\zeta(s)}$ by the variant $\frac{\zeta''(s)}{\zeta(s)}$. Assuming for simplicity that all zeroes are simple, this function still has a simple pole at every zero $\rho$ (with residue $\frac{\zeta''(\rho)}{\zeta'(\rho)}$), but now has a <em>double</em> pole at $s=1$ (with residue $2$). As it turns out, this variant also has a Dirichlet series representation $$ \sum_n \frac{ \Lambda_2(n) }{n^s} = \frac{\zeta''(s)}{\zeta(s)}$$ which by Perron's formula suggests an asymptotic of the form $$ \sum_{n \leq x} \Lambda_2(n) = 2 x \log x + \sum_\rho \frac{\zeta''(\rho)}{\zeta'(\rho)} \frac{x^\rho}{\rho} + \dots$$ Now, even if there are zeroes on $\mathrm{Re}(s)=1$, their contribution should still be smaller than the main term, and one may now be led to predict an asymptotic of the form $$ \sum_{n \leq x} \Lambda_2(n) = 2 x \log x + O(x) \quad (4)$$ and furthermore this asymptotic should be easier to prove than the prime number theorem, as it is not reliant on preventing zeroes at the edge of the critical strip.</p> <p>The functions $\Lambda_2$ and $\Lambda$ are related via the identity $$ \frac{\zeta''(s)}{\zeta(s)} = -\frac{d}{ds} (-\frac{\zeta'(s)}{\zeta(s)}) + (-\frac{\zeta'(s)}{\zeta(s)})^2$$ which after inverting the Dirichlet series leads to $$ \Lambda_2(n) = \Lambda(n) \log n + \sum_{d|n} \Lambda(d) \Lambda(\frac{n}{d}) \quad (5)$$ and from this and (4) we soon obtain (2) and hence (1).</p> <p>One can also think about this from a "music of the primes" standpoint. The explicit formula is morally resolving $\Lambda$ into "harmonics" as $$ \Lambda(n) \approx 1 - \sum_\rho n^{\rho-1} + \dots \quad (6)$$ where I am deliberately being vague as to how to interpret the symbols $\approx$, $\sum$, and $\dots$. The Dirichlet convolution of $1$ with itself resembles $\log n$, while the Dirichlet convolution of $-n^{\rho-1}$ with itself resembles $n^{\rho-1} \log n$; also, cross-terms such as the convolution of $1$ with $-n^{\rho-1}$ experience significant cancellation and should be lower order. Thus the right-hand side of (5) amplifies the "1" harmonic of $\Lambda$ by a factor of about $2 \log n$, while damping out the contributions of each of the zeroes $\rho$ due to cancellation between the two terms, leading one to (1) or (2) as a smoothed out version of the prime number theorem. Conversely, (2) combined with (5) can be viewed as an elementary way to encode (the largest terms of) the explicit formula (4), without explicit mention of zeroes.</p> <p>The most difficult part of the Erdos-Selberg elementary proof of the PNT is the passage from (1) to the prime number theorem, as here we must actually eliminate the scenario in which there is a zero on $\mathrm{Re}(s)=1$. The key to doing so is to exploit the non-negativity of $\Lambda$; morally, the expansion (6) prevents $\rho$ (and hence also $\overline{\rho}$) from lying on $\mathrm{Re}(s)=1$ as this would make the right-hand side RHS negative "on average" for certain sets of $n$. There are several ways to make this intuition precise; the standard complex-analytic proofs often proceed by testing $\Lambda$ against $n^{\pm it}$ and $n^{\pm 2it}$, where $1+it$ is a hypothetical zero of $\zeta$. Very roughly speaking, the idea of the elementary proof is to exploit not only the non-negativity of $\Lambda$, but also upper bounds coming from sieve-theoretic bounds such as the <a href="https://en.wikipedia.org/wiki/Brun%E2%80%93Titchmarsh_theorem" rel="noreferrer">Brun-Titchmarsh inequality</a>, which roughly speaking tell us that $\Lambda$ behaves "on average" as if it lies between $0$ and $2$. (Actually, the Selberg formula (2) already gives an adequate substitute for this inequality for the purposes of this argument.) On the other hand, from (6) we see that if there was a zero $\rho = 1+it$ on the line $\mathrm{Re}(s)=1$, then $\Lambda-1$ should have an extremely large correlation with $\mathrm{Re} n^{\rho-1} = \cos(t \log n)$, and these two facts can be placed in contradiction with each other, basically because the average value of $|\cos(t \log n)|$ is strictly less than one. However, because the zeroes $\rho$ are not explicitly present in the Selberg formula (2), one has to work quite a bit in the elementary proof to make this argument rigorous. In <a href="https://terrytao.wordpress.com/2014/10/25/a-banach-algebra-proof-of-the-prime-number-theorem/" rel="noreferrer">this blog post of mine</a> I use some Banach algebra theory to tease out the zeroes $\rho=1+it$ from the Selberg formula (they can be interpreted as the "Gelfand spectrum" of a certain Banach algebra norm associated to that formula) to make this connection more explicit.</p> Answer 2: <p>Complementing Terry's nice response, let me try to explain from a more elementary point of view why Selberg's formula (1) is natural, and why it is true.</p> <p>It is natural to formulate the PNT in terms of the von Mangoldt function $\Lambda(n)$, because it is supported on the prime powers, where it is given by the simple analytic formula $\Lambda(p^r)=\log p$. In addition, $\Lambda(n)$ is intimately related to the Möbius function $\mu(n)$ and the Riemann-zeta function via the equivalent formulae $$ \Lambda=\mu\ast\log,\qquad 1\ast\Lambda=\log,\qquad \sum_{n=1}^\infty\frac{\Lambda(n)}{n^s}=-\frac{\zeta'}{\zeta}(s).$$ Selberg's insight was that a natural and slightly more accessible quantity arises if we convolve $\mu$ not with $\log$, but a power of $\log$: $$ \Lambda_k=\mu\ast\log^k,\qquad 1\ast\Lambda_k=\log^k,\qquad \sum_{n=1}^\infty\frac{\Lambda_k(n)}{n^s}=(-1)^k\frac{\zeta^{(k)}}{\zeta}(s).$$ It is a nice exercise that $\Lambda_k(n)$ is nonnegative, hence at most $\log^k(n)$, and it is supported on numbers with at most $k$ distinct prime factors. Now (1) is essentially the same as $$ \sum_{n\leq x}\Lambda_2(n)=2x\log x+O(x), $$ i.e. an analogue of the PNT for numbers with at most two distinct prime factors (which makes the sum smoother). This explains why (1) is so natural and useful to look at. It is another matter, and the ingenuity of Selberg and Erdős, that one can derive information from the distribution of numbers with at most two distinct prime factors for the distribution of prime powers.</p> <p>In general, we have the following generalization of Selberg's formula (1): $$ \psi_k(x):=\sum_{n\leq x}\Lambda_k(n)=x P_k(\log x)+O(x),\qquad k\geq 2,$$ where $P_k(t)\in\mathbb{R}[t]$ is a polynomial of degree $k-1$ with leading coefficient $k$. Let me sketch a proof of this more general formula. First, using the definitions, $$ \psi_k(x)=\sum_{n\leq x}\mu(n)L_k(x/n),$$ where $L_k$ is the summatory function of $\log^k$: $$ L_k(x):=\sum_{n\leq x}\log^k(n).$$ A key point is that $L_k(x)$ is a smooth sum, in fact by simple calculus $$ L_k(x)=xR_k(\log x)+O(\log^k(x)), $$ where $R_k(t)\in\mathbb{R}[t]$ is a polynomial of degree $k$ with leading coefficient $1$. Now the idea is to compare $L_k$ with the summatory functions of the generalized divisor functions: $$ T_j(x):=\sum_{n\leq x}\tau_j(n),\qquad \tau_j:=1\ast\dots\ast 1. $$ These are easier to analyze than $\psi_k(x)$, in fact one can prove in an elementary way that $$ T_j(x)=xS_j(\log x)+O(x^{1-1/j}), \qquad j\geq 1,$$ where $S_j(t)\in\mathbb{R}[t]$ is a polynomial of degree $j-1$ with leading coefficient $1/(j-1)!$. Here we used the convention that $\tau_1:=1$, so that $T_1(x)=\lfloor x\rfloor=x+O(1)$ as claimed. It follows that $L_k(x)$ can be well-approximated with a linear combination of the $T_j(x)$'s: $$ L_k(x)=k!T_{k+1}(x)+a_kT_k(x)+\dots+a_1T_1(x)+O(x^{1-1/k}).$$ Plugging this back into our formula for $\psi_k$, we get $$ \psi_k(x)=\sum_{n\leq x}\mu(n)\Bigl\{k!T_{k+1}(x/n)+a_kT_k(x/n)+\dots+a_1T_1(x/n)\Bigr\}+O(x).$$ However, for each $j\geq 2$, $$ \sum_{n\leq x}\mu(n)T_j(x/n)=\sum_{n\leq x}(\mu\ast\tau_j)(n)=\sum_{n\leq x}\tau_{j-1}(n)=T_{j-1}(x), $$ and also $$ \sum_{n\leq x}\mu(n)T_1(x/n)=\sum_{n\leq x}(\mu\ast 1)(n)=1, $$ hence in fact $$\begin{align*} \psi_k(x)&amp;=k!T_k(x)+a_kT_{k-1}(x)+\dots +a_2 T_1(x)+a_1+O(x)\\ &amp;=x\Bigl\{k!S_k(\log x)+a_kS_{k-1}(\log x)+\dots + a_2 S_1(\log x)\Bigr\} + O(x).\end{align*} $$ This proves the claimed approximation for $\psi_k(x)$ with the polynomial $$P_k(t):=k!S_k(t)+a_kS_{k-1}(t)+\dots+a_2 S_1(t).$$ This polynomial has real coefficients, degree $k-1$ and leading coefficient $k!/(k-1)!=k$, where we used that $k\geq 2$ throughout. The proof is complete.</p> Answer 3: <p>Two very nice answers have already been given, but I would like to add that Michel Balazard has a book in preparation on this topic (written for undergrads), in which he tries to give a deduction of the PNT from Selberg's identity which as simple as possible. Also, a student here has a nice <a href="https://www.eleves.ens.fr/home/guignard/pdf/pnt.pdf" rel="nofollow noreferrer">write-up</a> of the proof. It is very short, it is elementary, but it is in french.</p> Answer 4: <p>The complex-analytic proof of the prime number theorem can help inform the elementary one.</p> <p>The von Mangoldt function $\Lambda$ is related to the Riemann zeta function $\zeta$ by the formula $$ \sum_n \frac{\Lambda(n)}{n^s} = -\frac{\zeta'(s)}{\zeta(s)},$$ at least in the region $\mathrm{Re}(s) &gt; 1$. The right-hand side has a pole of residue 1 at the simple pole $s=1$ of $\zeta$, and a pole of residue -1 at each zero $s=\rho$ of $\zeta$. Applying Perron's formula carefully, one eventually arrives at the von Mangoldt <a href="https://en.wikipedia.org/wiki/Explicit_formulae_(L-function)" rel="noreferrer">explicit formula</a></p> <p>$$ \sum_{n \leq x} \Lambda(n) = x - \sum_\rho \frac{x^\rho}{\rho} + \dots$$ where the remaining terms $\dots$ (as well as the way in which the infinite sum over zeroes is interpreted) will be ignored for this discussion. This formula strongly suggests that zeroes $\rho$ with real part strictly less than one will eventually only make a lower order contribution to $\sum_{n \leq x} \Lambda(n)$, whereas zeroes $\rho$ with real part equal to one would likely destroy the prime number theorem. In fact one can pursue this line of reasoning more carefully and show that the prime number theorem is equivalent to the absence of zeroes of zeta on the line $\mathrm{Re}(s)=1$.</p> <p>But what if there was a way to make the contribution of the zeroes of zeta smaller than those of the pole, even when said zeroes lie on the line $\mathrm{Re}(s)=1$? One way to do this is to replace the log-derivative $-\frac{\zeta'(s)}{\zeta(s)}$ by the variant $\frac{\zeta''(s)}{\zeta(s)}$. Assuming for simplicity that all zeroes are simple, this function still has a simple pole at every zero $\rho$ (with residue $\frac{\zeta''(\rho)}{\zeta'(\rho)}$), but now has a <em>double</em> pole at $s=1$ (with residue $2$). As it turns out, this variant also has a Dirichlet series representation $$ \sum_n \frac{ \Lambda_2(n) }{n^s} = \frac{\zeta''(s)}{\zeta(s)}$$ which by Perron's formula suggests an asymptotic of the form $$ \sum_{n \leq x} \Lambda_2(n) = 2 x \log x + \sum_\rho \frac{\zeta''(\rho)}{\zeta'(\rho)} \frac{x^\rho}{\rho} + \dots$$ Now, even if there are zeroes on $\mathrm{Re}(s)=1$, their contribution should still be smaller than the main term, and one may now be led to predict an asymptotic of the form $$ \sum_{n \leq x} \Lambda_2(n) = 2 x \log x + O(x) \quad (4)$$ and furthermore this asymptotic should be easier to prove than the prime number theorem, as it is not reliant on preventing zeroes at the edge of the critical strip.</p> <p>The functions $\Lambda_2$ and $\Lambda$ are related via the identity $$ \frac{\zeta''(s)}{\zeta(s)} = -\frac{d}{ds} (-\frac{\zeta'(s)}{\zeta(s)}) + (-\frac{\zeta'(s)}{\zeta(s)})^2$$ which after inverting the Dirichlet series leads to $$ \Lambda_2(n) = \Lambda(n) \log n + \sum_{d|n} \Lambda(d) \Lambda(\frac{n}{d}) \quad (5)$$ and from this and (4) we soon obtain (2) and hence (1).</p> <p>One can also think about this from a "music of the primes" standpoint. The explicit formula is morally resolving $\Lambda$ into "harmonics" as $$ \Lambda(n) \approx 1 - \sum_\rho n^{\rho-1} + \dots \quad (6)$$ where I am deliberately being vague as to how to interpret the symbols $\approx$, $\sum$, and $\dots$. The Dirichlet convolution of $1$ with itself resembles $\log n$, while the Dirichlet convolution of $-n^{\rho-1}$ with itself resembles $n^{\rho-1} \log n$; also, cross-terms such as the convolution of $1$ with $-n^{\rho-1}$ experience significant cancellation and should be lower order. Thus the right-hand side of (5) amplifies the "1" harmonic of $\Lambda$ by a factor of about $2 \log n$, while damping out the contributions of each of the zeroes $\rho$ due to cancellation between the two terms, leading one to (1) or (2) as a smoothed out version of the prime number theorem. Conversely, (2) combined with (5) can be viewed as an elementary way to encode (the largest terms of) the explicit formula (4), without explicit mention of zeroes.</p> <p>The most difficult part of the Erdos-Selberg elementary proof of the PNT is the passage from (1) to the prime number theorem, as here we must actually eliminate the scenario in which there is a zero on $\mathrm{Re}(s)=1$. The key to doing so is to exploit the non-negativity of $\Lambda$; morally, the expansion (6) prevents $\rho$ (and hence also $\overline{\rho}$) from lying on $\mathrm{Re}(s)=1$ as this would make the right-hand side RHS negative "on average" for certain sets of $n$. There are several ways to make this intuition precise; the standard complex-analytic proofs often proceed by testing $\Lambda$ against $n^{\pm it}$ and $n^{\pm 2it}$, where $1+it$ is a hypothetical zero of $\zeta$. Very roughly speaking, the idea of the elementary proof is to exploit not only the non-negativity of $\Lambda$, but also upper bounds coming from sieve-theoretic bounds such as the <a href="https://en.wikipedia.org/wiki/Brun%E2%80%93Titchmarsh_theorem" rel="noreferrer">Brun-Titchmarsh inequality</a>, which roughly speaking tell us that $\Lambda$ behaves "on average" as if it lies between $0$ and $2$. (Actually, the Selberg formula (2) already gives an adequate substitute for this inequality for the purposes of this argument.) On the other hand, from (6) we see that if there was a zero $\rho = 1+it$ on the line $\mathrm{Re}(s)=1$, then $\Lambda-1$ should have an extremely large correlation with $\mathrm{Re} n^{\rho-1} = \cos(t \log n)$, and these two facts can be placed in contradiction with each other, basically because the average value of $|\cos(t \log n)|$ is strictly less than one. However, because the zeroes $\rho$ are not explicitly present in the Selberg formula (2), one has to work quite a bit in the elementary proof to make this argument rigorous. In <a href="https://terrytao.wordpress.com/2014/10/25/a-banach-algebra-proof-of-the-prime-number-theorem/" rel="noreferrer">this blog post of mine</a> I use some Banach algebra theory to tease out the zeroes $\rho=1+it$ from the Selberg formula (they can be interpreted as the "Gelfand spectrum" of a certain Banach algebra norm associated to that formula) to make this connection more explicit.</p> Answer 5: <p>Complementing Terry's nice response, let me try to explain from a more elementary point of view why Selberg's formula (1) is natural, and why it is true.</p> <p>It is natural to formulate the PNT in terms of the von Mangoldt function $\Lambda(n)$, because it is supported on the prime powers, where it is given by the simple analytic formula $\Lambda(p^r)=\log p$. In addition, $\Lambda(n)$ is intimately related to the Möbius function $\mu(n)$ and the Riemann-zeta function via the equivalent formulae $$ \Lambda=\mu\ast\log,\qquad 1\ast\Lambda=\log,\qquad \sum_{n=1}^\infty\frac{\Lambda(n)}{n^s}=-\frac{\zeta'}{\zeta}(s).$$ Selberg's insight was that a natural and slightly more accessible quantity arises if we convolve $\mu$ not with $\log$, but a power of $\log$: $$ \Lambda_k=\mu\ast\log^k,\qquad 1\ast\Lambda_k=\log^k,\qquad \sum_{n=1}^\infty\frac{\Lambda_k(n)}{n^s}=(-1)^k\frac{\zeta^{(k)}}{\zeta}(s).$$ It is a nice exercise that $\Lambda_k(n)$ is nonnegative, hence at most $\log^k(n)$, and it is supported on numbers with at most $k$ distinct prime factors. Now (1) is essentially the same as $$ \sum_{n\leq x}\Lambda_2(n)=2x\log x+O(x), $$ i.e. an analogue of the PNT for numbers with at most two distinct prime factors (which makes the sum smoother). This explains why (1) is so natural and useful to look at. It is another matter, and the ingenuity of Selberg and Erdős, that one can derive information from the distribution of numbers with at most two distinct prime factors for the distribution of prime powers.</p> <p>In general, we have the following generalization of Selberg's formula (1): $$ \psi_k(x):=\sum_{n\leq x}\Lambda_k(n)=x P_k(\log x)+O(x),\qquad k\geq 2,$$ where $P_k(t)\in\mathbb{R}[t]$ is a polynomial of degree $k-1$ with leading coefficient $k$. Let me sketch a proof of this more general formula. First, using the definitions, $$ \psi_k(x)=\sum_{n\leq x}\mu(n)L_k(x/n),$$ where $L_k$ is the summatory function of $\log^k$: $$ L_k(x):=\sum_{n\leq x}\log^k(n).$$ A key point is that $L_k(x)$ is a smooth sum, in fact by simple calculus $$ L_k(x)=xR_k(\log x)+O(\log^k(x)), $$ where $R_k(t)\in\mathbb{R}[t]$ is a polynomial of degree $k$ with leading coefficient $1$. Now the idea is to compare $L_k$ with the summatory functions of the generalized divisor functions: $$ T_j(x):=\sum_{n\leq x}\tau_j(n),\qquad \tau_j:=1\ast\dots\ast 1. $$ These are easier to analyze than $\psi_k(x)$, in fact one can prove in an elementary way that $$ T_j(x)=xS_j(\log x)+O(x^{1-1/j}), \qquad j\geq 1,$$ where $S_j(t)\in\mathbb{R}[t]$ is a polynomial of degree $j-1$ with leading coefficient $1/(j-1)!$. Here we used the convention that $\tau_1:=1$, so that $T_1(x)=\lfloor x\rfloor=x+O(1)$ as claimed. It follows that $L_k(x)$ can be well-approximated with a linear combination of the $T_j(x)$'s: $$ L_k(x)=k!T_{k+1}(x)+a_kT_k(x)+\dots+a_1T_1(x)+O(x^{1-1/k}).$$ Plugging this back into our formula for $\psi_k$, we get $$ \psi_k(x)=\sum_{n\leq x}\mu(n)\Bigl\{k!T_{k+1}(x/n)+a_kT_k(x/n)+\dots+a_1T_1(x/n)\Bigr\}+O(x).$$ However, for each $j\geq 2$, $$ \sum_{n\leq x}\mu(n)T_j(x/n)=\sum_{n\leq x}(\mu\ast\tau_j)(n)=\sum_{n\leq x}\tau_{j-1}(n)=T_{j-1}(x), $$ and also $$ \sum_{n\leq x}\mu(n)T_1(x/n)=\sum_{n\leq x}(\mu\ast 1)(n)=1, $$ hence in fact $$\begin{align*} \psi_k(x)&amp;=k!T_k(x)+a_kT_{k-1}(x)+\dots +a_2 T_1(x)+a_1+O(x)\\ &amp;=x\Bigl\{k!S_k(\log x)+a_kS_{k-1}(\log x)+\dots + a_2 S_1(\log x)\Bigr\} + O(x).\end{align*} $$ This proves the claimed approximation for $\psi_k(x)$ with the polynomial $$P_k(t):=k!S_k(t)+a_kS_{k-1}(t)+\dots+a_2 S_1(t).$$ This polynomial has real coefficients, degree $k-1$ and leading coefficient $k!/(k-1)!=k$, where we used that $k\geq 2$ throughout. The proof is complete.</p> Answer 6: <p>Two very nice answers have already been given, but I would like to add that Michel Balazard has a book in preparation on this topic (written for undergrads), in which he tries to give a deduction of the PNT from Selberg's identity which as simple as possible. Also, a student here has a nice <a href="https://www.eleves.ens.fr/home/guignard/pdf/pnt.pdf" rel="nofollow noreferrer">write-up</a> of the proof. It is very short, it is elementary, but it is in french.</p> Answer 7: <p>The complex-analytic proof of the prime number theorem can help inform the elementary one.</p> <p>The von Mangoldt function $\Lambda$ is related to the Riemann zeta function $\zeta$ by the formula $$ \sum_n \frac{\Lambda(n)}{n^s} = -\frac{\zeta'(s)}{\zeta(s)},$$ at least in the region $\mathrm{Re}(s) &gt; 1$. The right-hand side has a pole of residue 1 at the simple pole $s=1$ of $\zeta$, and a pole of residue -1 at each zero $s=\rho$ of $\zeta$. Applying Perron's formula carefully, one eventually arrives at the von Mangoldt <a href="https://en.wikipedia.org/wiki/Explicit_formulae_(L-function)" rel="noreferrer">explicit formula</a></p> <p>$$ \sum_{n \leq x} \Lambda(n) = x - \sum_\rho \frac{x^\rho}{\rho} + \dots$$ where the remaining terms $\dots$ (as well as the way in which the infinite sum over zeroes is interpreted) will be ignored for this discussion. This formula strongly suggests that zeroes $\rho$ with real part strictly less than one will eventually only make a lower order contribution to $\sum_{n \leq x} \Lambda(n)$, whereas zeroes $\rho$ with real part equal to one would likely destroy the prime number theorem. In fact one can pursue this line of reasoning more carefully and show that the prime number theorem is equivalent to the absence of zeroes of zeta on the line $\mathrm{Re}(s)=1$.</p> <p>But what if there was a way to make the contribution of the zeroes of zeta smaller than those of the pole, even when said zeroes lie on the line $\mathrm{Re}(s)=1$? One way to do this is to replace the log-derivative $-\frac{\zeta'(s)}{\zeta(s)}$ by the variant $\frac{\zeta''(s)}{\zeta(s)}$. Assuming for simplicity that all zeroes are simple, this function still has a simple pole at every zero $\rho$ (with residue $\frac{\zeta''(\rho)}{\zeta'(\rho)}$), but now has a <em>double</em> pole at $s=1$ (with residue $2$). As it turns out, this variant also has a Dirichlet series representation $$ \sum_n \frac{ \Lambda_2(n) }{n^s} = \frac{\zeta''(s)}{\zeta(s)}$$ which by Perron's formula suggests an asymptotic of the form $$ \sum_{n \leq x} \Lambda_2(n) = 2 x \log x + \sum_\rho \frac{\zeta''(\rho)}{\zeta'(\rho)} \frac{x^\rho}{\rho} + \dots$$ Now, even if there are zeroes on $\mathrm{Re}(s)=1$, their contribution should still be smaller than the main term, and one may now be led to predict an asymptotic of the form $$ \sum_{n \leq x} \Lambda_2(n) = 2 x \log x + O(x) \quad (4)$$ and furthermore this asymptotic should be easier to prove than the prime number theorem, as it is not reliant on preventing zeroes at the edge of the critical strip.</p> <p>The functions $\Lambda_2$ and $\Lambda$ are related via the identity $$ \frac{\zeta''(s)}{\zeta(s)} = -\frac{d}{ds} (-\frac{\zeta'(s)}{\zeta(s)}) + (-\frac{\zeta'(s)}{\zeta(s)})^2$$ which after inverting the Dirichlet series leads to $$ \Lambda_2(n) = \Lambda(n) \log n + \sum_{d|n} \Lambda(d) \Lambda(\frac{n}{d}) \quad (5)$$ and from this and (4) we soon obtain (2) and hence (1).</p> <p>One can also think about this from a "music of the primes" standpoint. The explicit formula is morally resolving $\Lambda$ into "harmonics" as $$ \Lambda(n) \approx 1 - \sum_\rho n^{\rho-1} + \dots \quad (6)$$ where I am deliberately being vague as to how to interpret the symbols $\approx$, $\sum$, and $\dots$. The Dirichlet convolution of $1$ with itself resembles $\log n$, while the Dirichlet convolution of $-n^{\rho-1}$ with itself resembles $n^{\rho-1} \log n$; also, cross-terms such as the convolution of $1$ with $-n^{\rho-1}$ experience significant cancellation and should be lower order. Thus the right-hand side of (5) amplifies the "1" harmonic of $\Lambda$ by a factor of about $2 \log n$, while damping out the contributions of each of the zeroes $\rho$ due to cancellation between the two terms, leading one to (1) or (2) as a smoothed out version of the prime number theorem. Conversely, (2) combined with (5) can be viewed as an elementary way to encode (the largest terms of) the explicit formula (4), without explicit mention of zeroes.</p> <p>The most difficult part of the Erdos-Selberg elementary proof of the PNT is the passage from (1) to the prime number theorem, as here we must actually eliminate the scenario in which there is a zero on $\mathrm{Re}(s)=1$. The key to doing so is to exploit the non-negativity of $\Lambda$; morally, the expansion (6) prevents $\rho$ (and hence also $\overline{\rho}$) from lying on $\mathrm{Re}(s)=1$ as this would make the right-hand side RHS negative "on average" for certain sets of $n$. There are several ways to make this intuition precise; the standard complex-analytic proofs often proceed by testing $\Lambda$ against $n^{\pm it}$ and $n^{\pm 2it}$, where $1+it$ is a hypothetical zero of $\zeta$. Very roughly speaking, the idea of the elementary proof is to exploit not only the non-negativity of $\Lambda$, but also upper bounds coming from sieve-theoretic bounds such as the <a href="https://en.wikipedia.org/wiki/Brun%E2%80%93Titchmarsh_theorem" rel="noreferrer">Brun-Titchmarsh inequality</a>, which roughly speaking tell us that $\Lambda$ behaves "on average" as if it lies between $0$ and $2$. (Actually, the Selberg formula (2) already gives an adequate substitute for this inequality for the purposes of this argument.) On the other hand, from (6) we see that if there was a zero $\rho = 1+it$ on the line $\mathrm{Re}(s)=1$, then $\Lambda-1$ should have an extremely large correlation with $\mathrm{Re} n^{\rho-1} = \cos(t \log n)$, and these two facts can be placed in contradiction with each other, basically because the average value of $|\cos(t \log n)|$ is strictly less than one. However, because the zeroes $\rho$ are not explicitly present in the Selberg formula (2), one has to work quite a bit in the elementary proof to make this argument rigorous. In <a href="https://terrytao.wordpress.com/2014/10/25/a-banach-algebra-proof-of-the-prime-number-theorem/" rel="noreferrer">this blog post of mine</a> I use some Banach algebra theory to tease out the zeroes $\rho=1+it$ from the Selberg formula (they can be interpreted as the "Gelfand spectrum" of a certain Banach algebra norm associated to that formula) to make this connection more explicit.</p> Answer 8: <p>Complementing Terry's nice response, let me try to explain from a more elementary point of view why Selberg's formula (1) is natural, and why it is true.</p> <p>It is natural to formulate the PNT in terms of the von Mangoldt function $\Lambda(n)$, because it is supported on the prime powers, where it is given by the simple analytic formula $\Lambda(p^r)=\log p$. In addition, $\Lambda(n)$ is intimately related to the Möbius function $\mu(n)$ and the Riemann-zeta function via the equivalent formulae $$ \Lambda=\mu\ast\log,\qquad 1\ast\Lambda=\log,\qquad \sum_{n=1}^\infty\frac{\Lambda(n)}{n^s}=-\frac{\zeta'}{\zeta}(s).$$ Selberg's insight was that a natural and slightly more accessible quantity arises if we convolve $\mu$ not with $\log$, but a power of $\log$: $$ \Lambda_k=\mu\ast\log^k,\qquad 1\ast\Lambda_k=\log^k,\qquad \sum_{n=1}^\infty\frac{\Lambda_k(n)}{n^s}=(-1)^k\frac{\zeta^{(k)}}{\zeta}(s).$$ It is a nice exercise that $\Lambda_k(n)$ is nonnegative, hence at most $\log^k(n)$, and it is supported on numbers with at most $k$ distinct prime factors. Now (1) is essentially the same as $$ \sum_{n\leq x}\Lambda_2(n)=2x\log x+O(x), $$ i.e. an analogue of the PNT for numbers with at most two distinct prime factors (which makes the sum smoother). This explains why (1) is so natural and useful to look at. It is another matter, and the ingenuity of Selberg and Erdős, that one can derive information from the distribution of numbers with at most two distinct prime factors for the distribution of prime powers.</p> <p>In general, we have the following generalization of Selberg's formula (1): $$ \psi_k(x):=\sum_{n\leq x}\Lambda_k(n)=x P_k(\log x)+O(x),\qquad k\geq 2,$$ where $P_k(t)\in\mathbb{R}[t]$ is a polynomial of degree $k-1$ with leading coefficient $k$. Let me sketch a proof of this more general formula. First, using the definitions, $$ \psi_k(x)=\sum_{n\leq x}\mu(n)L_k(x/n),$$ where $L_k$ is the summatory function of $\log^k$: $$ L_k(x):=\sum_{n\leq x}\log^k(n).$$ A key point is that $L_k(x)$ is a smooth sum, in fact by simple calculus $$ L_k(x)=xR_k(\log x)+O(\log^k(x)), $$ where $R_k(t)\in\mathbb{R}[t]$ is a polynomial of degree $k$ with leading coefficient $1$. Now the idea is to compare $L_k$ with the summatory functions of the generalized divisor functions: $$ T_j(x):=\sum_{n\leq x}\tau_j(n),\qquad \tau_j:=1\ast\dots\ast 1. $$ These are easier to analyze than $\psi_k(x)$, in fact one can prove in an elementary way that $$ T_j(x)=xS_j(\log x)+O(x^{1-1/j}), \qquad j\geq 1,$$ where $S_j(t)\in\mathbb{R}[t]$ is a polynomial of degree $j-1$ with leading coefficient $1/(j-1)!$. Here we used the convention that $\tau_1:=1$, so that $T_1(x)=\lfloor x\rfloor=x+O(1)$ as claimed. It follows that $L_k(x)$ can be well-approximated with a linear combination of the $T_j(x)$'s: $$ L_k(x)=k!T_{k+1}(x)+a_kT_k(x)+\dots+a_1T_1(x)+O(x^{1-1/k}).$$ Plugging this back into our formula for $\psi_k$, we get $$ \psi_k(x)=\sum_{n\leq x}\mu(n)\Bigl\{k!T_{k+1}(x/n)+a_kT_k(x/n)+\dots+a_1T_1(x/n)\Bigr\}+O(x).$$ However, for each $j\geq 2$, $$ \sum_{n\leq x}\mu(n)T_j(x/n)=\sum_{n\leq x}(\mu\ast\tau_j)(n)=\sum_{n\leq x}\tau_{j-1}(n)=T_{j-1}(x), $$ and also $$ \sum_{n\leq x}\mu(n)T_1(x/n)=\sum_{n\leq x}(\mu\ast 1)(n)=1, $$ hence in fact $$\begin{align*} \psi_k(x)&amp;=k!T_k(x)+a_kT_{k-1}(x)+\dots +a_2 T_1(x)+a_1+O(x)\\ &amp;=x\Bigl\{k!S_k(\log x)+a_kS_{k-1}(\log x)+\dots + a_2 S_1(\log x)\Bigr\} + O(x).\end{align*} $$ This proves the claimed approximation for $\psi_k(x)$ with the polynomial $$P_k(t):=k!S_k(t)+a_kS_{k-1}(t)+\dots+a_2 S_1(t).$$ This polynomial has real coefficients, degree $k-1$ and leading coefficient $k!/(k-1)!=k$, where we used that $k\geq 2$ throughout. The proof is complete.</p> Answer 9: <p>Two very nice answers have already been given, but I would like to add that Michel Balazard has a book in preparation on this topic (written for undergrads), in which he tries to give a deduction of the PNT from Selberg's identity which as simple as possible. Also, a student here has a nice <a href="https://www.eleves.ens.fr/home/guignard/pdf/pnt.pdf" rel="nofollow noreferrer">write-up</a> of the proof. It is very short, it is elementary, but it is in french.</p> Answer 10: <p>The complex-analytic proof of the prime number theorem can help inform the elementary one.</p> <p>The von Mangoldt function $\Lambda$ is related to the Riemann zeta function $\zeta$ by the formula $$ \sum_n \frac{\Lambda(n)}{n^s} = -\frac{\zeta'(s)}{\zeta(s)},$$ at least in the region $\mathrm{Re}(s) &gt; 1$. The right-hand side has a pole of residue 1 at the simple pole $s=1$ of $\zeta$, and a pole of residue -1 at each zero $s=\rho$ of $\zeta$. Applying Perron's formula carefully, one eventually arrives at the von Mangoldt <a href="https://en.wikipedia.org/wiki/Explicit_formulae_(L-function)" rel="noreferrer">explicit formula</a></p> <p>$$ \sum_{n \leq x} \Lambda(n) = x - \sum_\rho \frac{x^\rho}{\rho} + \dots$$ where the remaining terms $\dots$ (as well as the way in which the infinite sum over zeroes is interpreted) will be ignored for this discussion. This formula strongly suggests that zeroes $\rho$ with real part strictly less than one will eventually only make a lower order contribution to $\sum_{n \leq x} \Lambda(n)$, whereas zeroes $\rho$ with real part equal to one would likely destroy the prime number theorem. In fact one can pursue this line of reasoning more carefully and show that the prime number theorem is equivalent to the absence of zeroes of zeta on the line $\mathrm{Re}(s)=1$.</p> <p>But what if there was a way to make the contribution of the zeroes of zeta smaller than those of the pole, even when said zeroes lie on the line $\mathrm{Re}(s)=1$? One way to do this is to replace the log-derivative $-\frac{\zeta'(s)}{\zeta(s)}$ by the variant $\frac{\zeta''(s)}{\zeta(s)}$. Assuming for simplicity that all zeroes are simple, this function still has a simple pole at every zero $\rho$ (with residue $\frac{\zeta''(\rho)}{\zeta'(\rho)}$), but now has a <em>double</em> pole at $s=1$ (with residue $2$). As it turns out, this variant also has a Dirichlet series representation $$ \sum_n \frac{ \Lambda_2(n) }{n^s} = \frac{\zeta''(s)}{\zeta(s)}$$ which by Perron's formula suggests an asymptotic of the form $$ \sum_{n \leq x} \Lambda_2(n) = 2 x \log x + \sum_\rho \frac{\zeta''(\rho)}{\zeta'(\rho)} \frac{x^\rho}{\rho} + \dots$$ Now, even if there are zeroes on $\mathrm{Re}(s)=1$, their contribution should still be smaller than the main term, and one may now be led to predict an asymptotic of the form $$ \sum_{n \leq x} \Lambda_2(n) = 2 x \log x + O(x) \quad (4)$$ and furthermore this asymptotic should be easier to prove than the prime number theorem, as it is not reliant on preventing zeroes at the edge of the critical strip.</p> <p>The functions $\Lambda_2$ and $\Lambda$ are related via the identity $$ \frac{\zeta''(s)}{\zeta(s)} = -\frac{d}{ds} (-\frac{\zeta'(s)}{\zeta(s)}) + (-\frac{\zeta'(s)}{\zeta(s)})^2$$ which after inverting the Dirichlet series leads to $$ \Lambda_2(n) = \Lambda(n) \log n + \sum_{d|n} \Lambda(d) \Lambda(\frac{n}{d}) \quad (5)$$ and from this and (4) we soon obtain (2) and hence (1).</p> <p>One can also think about this from a "music of the primes" standpoint. The explicit formula is morally resolving $\Lambda$ into "harmonics" as $$ \Lambda(n) \approx 1 - \sum_\rho n^{\rho-1} + \dots \quad (6)$$ where I am deliberately being vague as to how to interpret the symbols $\approx$, $\sum$, and $\dots$. The Dirichlet convolution of $1$ with itself resembles $\log n$, while the Dirichlet convolution of $-n^{\rho-1}$ with itself resembles $n^{\rho-1} \log n$; also, cross-terms such as the convolution of $1$ with $-n^{\rho-1}$ experience significant cancellation and should be lower order. Thus the right-hand side of (5) amplifies the "1" harmonic of $\Lambda$ by a factor of about $2 \log n$, while damping out the contributions of each of the zeroes $\rho$ due to cancellation between the two terms, leading one to (1) or (2) as a smoothed out version of the prime number theorem. Conversely, (2) combined with (5) can be viewed as an elementary way to encode (the largest terms of) the explicit formula (4), without explicit mention of zeroes.</p> <p>The most difficult part of the Erdos-Selberg elementary proof of the PNT is the passage from (1) to the prime number theorem, as here we must actually eliminate the scenario in which there is a zero on $\mathrm{Re}(s)=1$. The key to doing so is to exploit the non-negativity of $\Lambda$; morally, the expansion (6) prevents $\rho$ (and hence also $\overline{\rho}$) from lying on $\mathrm{Re}(s)=1$ as this would make the right-hand side RHS negative "on average" for certain sets of $n$. There are several ways to make this intuition precise; the standard complex-analytic proofs often proceed by testing $\Lambda$ against $n^{\pm it}$ and $n^{\pm 2it}$, where $1+it$ is a hypothetical zero of $\zeta$. Very roughly speaking, the idea of the elementary proof is to exploit not only the non-negativity of $\Lambda$, but also upper bounds coming from sieve-theoretic bounds such as the <a href="https://en.wikipedia.org/wiki/Brun%E2%80%93Titchmarsh_theorem" rel="noreferrer">Brun-Titchmarsh inequality</a>, which roughly speaking tell us that $\Lambda$ behaves "on average" as if it lies between $0$ and $2$. (Actually, the Selberg formula (2) already gives an adequate substitute for this inequality for the purposes of this argument.) On the other hand, from (6) we see that if there was a zero $\rho = 1+it$ on the line $\mathrm{Re}(s)=1$, then $\Lambda-1$ should have an extremely large correlation with $\mathrm{Re} n^{\rho-1} = \cos(t \log n)$, and these two facts can be placed in contradiction with each other, basically because the average value of $|\cos(t \log n)|$ is strictly less than one. However, because the zeroes $\rho$ are not explicitly present in the Selberg formula (2), one has to work quite a bit in the elementary proof to make this argument rigorous. In <a href="https://terrytao.wordpress.com/2014/10/25/a-banach-algebra-proof-of-the-prime-number-theorem/" rel="noreferrer">this blog post of mine</a> I use some Banach algebra theory to tease out the zeroes $\rho=1+it$ from the Selberg formula (they can be interpreted as the "Gelfand spectrum" of a certain Banach algebra norm associated to that formula) to make this connection more explicit.</p> Answer 11: <p>Complementing Terry's nice response, let me try to explain from a more elementary point of view why Selberg's formula (1) is natural, and why it is true.</p> <p>It is natural to formulate the PNT in terms of the von Mangoldt function $\Lambda(n)$, because it is supported on the prime powers, where it is given by the simple analytic formula $\Lambda(p^r)=\log p$. In addition, $\Lambda(n)$ is intimately related to the Möbius function $\mu(n)$ and the Riemann-zeta function via the equivalent formulae $$ \Lambda=\mu\ast\log,\qquad 1\ast\Lambda=\log,\qquad \sum_{n=1}^\infty\frac{\Lambda(n)}{n^s}=-\frac{\zeta'}{\zeta}(s).$$ Selberg's insight was that a natural and slightly more accessible quantity arises if we convolve $\mu$ not with $\log$, but a power of $\log$: $$ \Lambda_k=\mu\ast\log^k,\qquad 1\ast\Lambda_k=\log^k,\qquad \sum_{n=1}^\infty\frac{\Lambda_k(n)}{n^s}=(-1)^k\frac{\zeta^{(k)}}{\zeta}(s).$$ It is a nice exercise that $\Lambda_k(n)$ is nonnegative, hence at most $\log^k(n)$, and it is supported on numbers with at most $k$ distinct prime factors. Now (1) is essentially the same as $$ \sum_{n\leq x}\Lambda_2(n)=2x\log x+O(x), $$ i.e. an analogue of the PNT for numbers with at most two distinct prime factors (which makes the sum smoother). This explains why (1) is so natural and useful to look at. It is another matter, and the ingenuity of Selberg and Erdős, that one can derive information from the distribution of numbers with at most two distinct prime factors for the distribution of prime powers.</p> <p>In general, we have the following generalization of Selberg's formula (1): $$ \psi_k(x):=\sum_{n\leq x}\Lambda_k(n)=x P_k(\log x)+O(x),\qquad k\geq 2,$$ where $P_k(t)\in\mathbb{R}[t]$ is a polynomial of degree $k-1$ with leading coefficient $k$. Let me sketch a proof of this more general formula. First, using the definitions, $$ \psi_k(x)=\sum_{n\leq x}\mu(n)L_k(x/n),$$ where $L_k$ is the summatory function of $\log^k$: $$ L_k(x):=\sum_{n\leq x}\log^k(n).$$ A key point is that $L_k(x)$ is a smooth sum, in fact by simple calculus $$ L_k(x)=xR_k(\log x)+O(\log^k(x)), $$ where $R_k(t)\in\mathbb{R}[t]$ is a polynomial of degree $k$ with leading coefficient $1$. Now the idea is to compare $L_k$ with the summatory functions of the generalized divisor functions: $$ T_j(x):=\sum_{n\leq x}\tau_j(n),\qquad \tau_j:=1\ast\dots\ast 1. $$ These are easier to analyze than $\psi_k(x)$, in fact one can prove in an elementary way that $$ T_j(x)=xS_j(\log x)+O(x^{1-1/j}), \qquad j\geq 1,$$ where $S_j(t)\in\mathbb{R}[t]$ is a polynomial of degree $j-1$ with leading coefficient $1/(j-1)!$. Here we used the convention that $\tau_1:=1$, so that $T_1(x)=\lfloor x\rfloor=x+O(1)$ as claimed. It follows that $L_k(x)$ can be well-approximated with a linear combination of the $T_j(x)$'s: $$ L_k(x)=k!T_{k+1}(x)+a_kT_k(x)+\dots+a_1T_1(x)+O(x^{1-1/k}).$$ Plugging this back into our formula for $\psi_k$, we get $$ \psi_k(x)=\sum_{n\leq x}\mu(n)\Bigl\{k!T_{k+1}(x/n)+a_kT_k(x/n)+\dots+a_1T_1(x/n)\Bigr\}+O(x).$$ However, for each $j\geq 2$, $$ \sum_{n\leq x}\mu(n)T_j(x/n)=\sum_{n\leq x}(\mu\ast\tau_j)(n)=\sum_{n\leq x}\tau_{j-1}(n)=T_{j-1}(x), $$ and also $$ \sum_{n\leq x}\mu(n)T_1(x/n)=\sum_{n\leq x}(\mu\ast 1)(n)=1, $$ hence in fact $$\begin{align*} \psi_k(x)&amp;=k!T_k(x)+a_kT_{k-1}(x)+\dots +a_2 T_1(x)+a_1+O(x)\\ &amp;=x\Bigl\{k!S_k(\log x)+a_kS_{k-1}(\log x)+\dots + a_2 S_1(\log x)\Bigr\} + O(x).\end{align*} $$ This proves the claimed approximation for $\psi_k(x)$ with the polynomial $$P_k(t):=k!S_k(t)+a_kS_{k-1}(t)+\dots+a_2 S_1(t).$$ This polynomial has real coefficients, degree $k-1$ and leading coefficient $k!/(k-1)!=k$, where we used that $k\geq 2$ throughout. The proof is complete.</p> Answer 12: <p>Two very nice answers have already been given, but I would like to add that Michel Balazard has a book in preparation on this topic (written for undergrads), in which he tries to give a deduction of the PNT from Selberg's identity which as simple as possible. Also, a student here has a nice <a href="https://www.eleves.ens.fr/home/guignard/pdf/pnt.pdf" rel="nofollow noreferrer">write-up</a> of the proof. It is very short, it is elementary, but it is in french.</p>
Why are Fuchsian groups interesting?
260,193
30
https://mathoverflow.net/questions/260193/why-are-fuchsian-groups-interesting
<p>I keep hearing that fuchsian groups are interesting for other reasons than the Fuchsian model for hyperbolic Riemann surfaces. </p> <p>What are those reasons? </p> <p>Are the Fuchsian groups with fixed points interesting from a geometric perspective?</p> <p>Where do Fuchsian groups appear besides hyperbolic geometry? </p> <p>I also read somewhere about a relation between fuchsian groups and fractals. Does someone know more about that and/or has a good reference? </p>
[ { "answer_id": 260199, "score": 21, "text": "<p>About the relation to fractals: for Fuchsian groups of the first kind, the limit set has Hausdorff dimension 1, i.e., it is not fractal. </p>\n\n<p>However, for all other quasifuchsian groups of the first kind, the limit set has Hausdorff dimension strictly bigger than 1, by a theorem of Rufus Bowen.</p>\n\n<p><a href=\"https://i.stack.imgur.com/WmL2K.gif\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/WmL2K.gif\" alt=\"enter image description here\"></a></p>\n" }, { "answer_id": 260194, "score": 18, "text": "<p>Check out <a href=\"http://rads.stackoverflow.com/amzn/click/1107564743\">Indra's Pearls.</a> (Mumford, Series, Wright).</p>\n" }, { "answer_id": 260258, "score": 13, "text": "<p>Fuchsian groups occur naturally in JSJ-theory. In 3-manifolds they occur as the base Groups of the Seifert pieces and in geometric group theory they occur as (quotients of) enclosing groups (also called QH-subgroups). In the case of hyperbolic groups these QH-subgroup carry in some sense the outer automorphism group of the whole group.</p>\n" }, { "answer_id": 260340, "score": 13, "text": "<p>Fuchsian groups, particularly those which are cocompact, form the tips of several big mathematical icebergs. To put this less metaphorically, several discoveries about Fuchsian groups, obtained using hyperbolic geometry, led to generalizations which fueled the growth of combinatorial and geometric group theory in the 20th century.</p>\n\n<p>By work of Dehn and others, cocompact Fuchsian groups are early examples of geometric computations of isoperimetric functions, aka Dehn functions, which Dehn discovered by working with group invariant polygonal tilings of the hyperbolic plane. </p>\n\n<p>Cocompact Fuchsian groups are also interesting early examples of groups with solvable word and conjugacy problem, using ideas related to Dehn's algorithm. </p>\n\n<p>Cocompact Fuchsian group (together with finite rank free groups) are the early examples of hyperbolic groups in the sense of Gromov, proved using that the group with its word metric is quasi-isometric to the hyperbolic plane.</p>\n" }, { "answer_id": 260239, "score": 11, "text": "<p>Fuchsian groups are important for the theory of <a href=\"https://en.wikipedia.org/wiki/Dessin_d%27enfant\" rel=\"nofollow noreferrer\">dessin d'enfants</a> in arithmetic geometry. Wolfart wrote a survey here: <a href=\"http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf\" rel=\"nofollow noreferrer\">http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf</a> .</p>\n\n<p>One typical application is the following: There is some $c&gt;0$, such that for infinitely many $g$ there are $&gt;g^{c\\log g}$ non-isomorphic complex curves $C$ of genus $g$ satisfying $|\\mathrm{Aut}(C)|=84(g-1)$. For the proof you connect the number of different curves to the number of normal subgroups of the $(2,3,7)$-triangle group, and then show that Fuchsian groups are virtually surface groups, and have therefore many normal subgroups.</p>\n" }, { "answer_id": 260372, "score": 7, "text": "<pre><code>Are the Fuchsian groups with fixed points interesting from a geometric perspective?\n</code></pre>\n\n<p>Yes, notice $PSL(2,\\mathbb{Z})$ fixes points. The fundamental domain in this case is isometric to a hyperbolic structure on the disk with two cone points. </p>\n\n<p>More generally, if a Fuchsian group $G$ has finite co-volume in $H^2$, then $H^2/G$ will be a hyperbolic 2-orbifold. For example, all of the orientation preserving subgroups of the hyperbolic triangle groups are Fuchsian groups of this type.</p>\n\n<p>In particular, hyperbolic triangle groups have a unique hyperbolic structure, so if 3-manifold has quotient which contains a (orientable) hyperbolic triangle orbifold then the original 3-manifold contains a totally geodesic surface. </p>\n" }, { "answer_id": 429141, "score": 0, "text": "<p>In Poincaré's words (in translation):</p>\n<blockquote>\n<p>The results so obtained as yet give only a very incomplete solution to\nthe problem I set myself, that is, the integration of linear\ndifferential equations. The equations I have called Fuchsian, and\nwhich can be integrated by means of a simple inversion, are just very\nspecial cases of second-order linear equations.</p>\n</blockquote>\n<p>See page 208 of <em>Uniformization of Riemman Surfaces</em> by Henri Paul de Saint-Gervais.</p>\n" }, { "answer_id": 260199, "score": 21, "text": "<p>About the relation to fractals: for Fuchsian groups of the first kind, the limit set has Hausdorff dimension 1, i.e., it is not fractal. </p>\n\n<p>However, for all other quasifuchsian groups of the first kind, the limit set has Hausdorff dimension strictly bigger than 1, by a theorem of Rufus Bowen.</p>\n\n<p><a href=\"https://i.stack.imgur.com/WmL2K.gif\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/WmL2K.gif\" alt=\"enter image description here\"></a></p>\n" }, { "answer_id": 260194, "score": 18, "text": "<p>Check out <a href=\"http://rads.stackoverflow.com/amzn/click/1107564743\">Indra's Pearls.</a> (Mumford, Series, Wright).</p>\n" }, { "answer_id": 260258, "score": 13, "text": "<p>Fuchsian groups occur naturally in JSJ-theory. In 3-manifolds they occur as the base Groups of the Seifert pieces and in geometric group theory they occur as (quotients of) enclosing groups (also called QH-subgroups). In the case of hyperbolic groups these QH-subgroup carry in some sense the outer automorphism group of the whole group.</p>\n" }, { "answer_id": 260340, "score": 13, "text": "<p>Fuchsian groups, particularly those which are cocompact, form the tips of several big mathematical icebergs. To put this less metaphorically, several discoveries about Fuchsian groups, obtained using hyperbolic geometry, led to generalizations which fueled the growth of combinatorial and geometric group theory in the 20th century.</p>\n\n<p>By work of Dehn and others, cocompact Fuchsian groups are early examples of geometric computations of isoperimetric functions, aka Dehn functions, which Dehn discovered by working with group invariant polygonal tilings of the hyperbolic plane. </p>\n\n<p>Cocompact Fuchsian groups are also interesting early examples of groups with solvable word and conjugacy problem, using ideas related to Dehn's algorithm. </p>\n\n<p>Cocompact Fuchsian group (together with finite rank free groups) are the early examples of hyperbolic groups in the sense of Gromov, proved using that the group with its word metric is quasi-isometric to the hyperbolic plane.</p>\n" }, { "answer_id": 260239, "score": 11, "text": "<p>Fuchsian groups are important for the theory of <a href=\"https://en.wikipedia.org/wiki/Dessin_d%27enfant\" rel=\"nofollow noreferrer\">dessin d'enfants</a> in arithmetic geometry. Wolfart wrote a survey here: <a href=\"http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf\" rel=\"nofollow noreferrer\">http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf</a> .</p>\n\n<p>One typical application is the following: There is some $c&gt;0$, such that for infinitely many $g$ there are $&gt;g^{c\\log g}$ non-isomorphic complex curves $C$ of genus $g$ satisfying $|\\mathrm{Aut}(C)|=84(g-1)$. For the proof you connect the number of different curves to the number of normal subgroups of the $(2,3,7)$-triangle group, and then show that Fuchsian groups are virtually surface groups, and have therefore many normal subgroups.</p>\n" }, { "answer_id": 260372, "score": 7, "text": "<pre><code>Are the Fuchsian groups with fixed points interesting from a geometric perspective?\n</code></pre>\n\n<p>Yes, notice $PSL(2,\\mathbb{Z})$ fixes points. The fundamental domain in this case is isometric to a hyperbolic structure on the disk with two cone points. </p>\n\n<p>More generally, if a Fuchsian group $G$ has finite co-volume in $H^2$, then $H^2/G$ will be a hyperbolic 2-orbifold. For example, all of the orientation preserving subgroups of the hyperbolic triangle groups are Fuchsian groups of this type.</p>\n\n<p>In particular, hyperbolic triangle groups have a unique hyperbolic structure, so if 3-manifold has quotient which contains a (orientable) hyperbolic triangle orbifold then the original 3-manifold contains a totally geodesic surface. </p>\n" }, { "answer_id": 429141, "score": 0, "text": "<p>In Poincaré's words (in translation):</p>\n<blockquote>\n<p>The results so obtained as yet give only a very incomplete solution to\nthe problem I set myself, that is, the integration of linear\ndifferential equations. The equations I have called Fuchsian, and\nwhich can be integrated by means of a simple inversion, are just very\nspecial cases of second-order linear equations.</p>\n</blockquote>\n<p>See page 208 of <em>Uniformization of Riemman Surfaces</em> by Henri Paul de Saint-Gervais.</p>\n" }, { "answer_id": 260199, "score": 21, "text": "<p>About the relation to fractals: for Fuchsian groups of the first kind, the limit set has Hausdorff dimension 1, i.e., it is not fractal. </p>\n\n<p>However, for all other quasifuchsian groups of the first kind, the limit set has Hausdorff dimension strictly bigger than 1, by a theorem of Rufus Bowen.</p>\n\n<p><a href=\"https://i.stack.imgur.com/WmL2K.gif\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/WmL2K.gif\" alt=\"enter image description here\"></a></p>\n" }, { "answer_id": 260194, "score": 18, "text": "<p>Check out <a href=\"http://rads.stackoverflow.com/amzn/click/1107564743\">Indra's Pearls.</a> (Mumford, Series, Wright).</p>\n" }, { "answer_id": 260258, "score": 13, "text": "<p>Fuchsian groups occur naturally in JSJ-theory. In 3-manifolds they occur as the base Groups of the Seifert pieces and in geometric group theory they occur as (quotients of) enclosing groups (also called QH-subgroups). In the case of hyperbolic groups these QH-subgroup carry in some sense the outer automorphism group of the whole group.</p>\n" }, { "answer_id": 260340, "score": 13, "text": "<p>Fuchsian groups, particularly those which are cocompact, form the tips of several big mathematical icebergs. To put this less metaphorically, several discoveries about Fuchsian groups, obtained using hyperbolic geometry, led to generalizations which fueled the growth of combinatorial and geometric group theory in the 20th century.</p>\n\n<p>By work of Dehn and others, cocompact Fuchsian groups are early examples of geometric computations of isoperimetric functions, aka Dehn functions, which Dehn discovered by working with group invariant polygonal tilings of the hyperbolic plane. </p>\n\n<p>Cocompact Fuchsian groups are also interesting early examples of groups with solvable word and conjugacy problem, using ideas related to Dehn's algorithm. </p>\n\n<p>Cocompact Fuchsian group (together with finite rank free groups) are the early examples of hyperbolic groups in the sense of Gromov, proved using that the group with its word metric is quasi-isometric to the hyperbolic plane.</p>\n" }, { "answer_id": 260239, "score": 11, "text": "<p>Fuchsian groups are important for the theory of <a href=\"https://en.wikipedia.org/wiki/Dessin_d%27enfant\" rel=\"nofollow noreferrer\">dessin d'enfants</a> in arithmetic geometry. Wolfart wrote a survey here: <a href=\"http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf\" rel=\"nofollow noreferrer\">http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf</a> .</p>\n\n<p>One typical application is the following: There is some $c&gt;0$, such that for infinitely many $g$ there are $&gt;g^{c\\log g}$ non-isomorphic complex curves $C$ of genus $g$ satisfying $|\\mathrm{Aut}(C)|=84(g-1)$. For the proof you connect the number of different curves to the number of normal subgroups of the $(2,3,7)$-triangle group, and then show that Fuchsian groups are virtually surface groups, and have therefore many normal subgroups.</p>\n" }, { "answer_id": 260372, "score": 7, "text": "<pre><code>Are the Fuchsian groups with fixed points interesting from a geometric perspective?\n</code></pre>\n\n<p>Yes, notice $PSL(2,\\mathbb{Z})$ fixes points. The fundamental domain in this case is isometric to a hyperbolic structure on the disk with two cone points. </p>\n\n<p>More generally, if a Fuchsian group $G$ has finite co-volume in $H^2$, then $H^2/G$ will be a hyperbolic 2-orbifold. For example, all of the orientation preserving subgroups of the hyperbolic triangle groups are Fuchsian groups of this type.</p>\n\n<p>In particular, hyperbolic triangle groups have a unique hyperbolic structure, so if 3-manifold has quotient which contains a (orientable) hyperbolic triangle orbifold then the original 3-manifold contains a totally geodesic surface. </p>\n" }, { "answer_id": 429141, "score": 0, "text": "<p>In Poincaré's words (in translation):</p>\n<blockquote>\n<p>The results so obtained as yet give only a very incomplete solution to\nthe problem I set myself, that is, the integration of linear\ndifferential equations. The equations I have called Fuchsian, and\nwhich can be integrated by means of a simple inversion, are just very\nspecial cases of second-order linear equations.</p>\n</blockquote>\n<p>See page 208 of <em>Uniformization of Riemman Surfaces</em> by Henri Paul de Saint-Gervais.</p>\n" }, { "answer_id": 260199, "score": 21, "text": "<p>About the relation to fractals: for Fuchsian groups of the first kind, the limit set has Hausdorff dimension 1, i.e., it is not fractal. </p>\n\n<p>However, for all other quasifuchsian groups of the first kind, the limit set has Hausdorff dimension strictly bigger than 1, by a theorem of Rufus Bowen.</p>\n\n<p><a href=\"https://i.stack.imgur.com/WmL2K.gif\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/WmL2K.gif\" alt=\"enter image description here\"></a></p>\n" }, { "answer_id": 260194, "score": 18, "text": "<p>Check out <a href=\"http://rads.stackoverflow.com/amzn/click/1107564743\">Indra's Pearls.</a> (Mumford, Series, Wright).</p>\n" }, { "answer_id": 260258, "score": 13, "text": "<p>Fuchsian groups occur naturally in JSJ-theory. In 3-manifolds they occur as the base Groups of the Seifert pieces and in geometric group theory they occur as (quotients of) enclosing groups (also called QH-subgroups). In the case of hyperbolic groups these QH-subgroup carry in some sense the outer automorphism group of the whole group.</p>\n" }, { "answer_id": 260340, "score": 13, "text": "<p>Fuchsian groups, particularly those which are cocompact, form the tips of several big mathematical icebergs. To put this less metaphorically, several discoveries about Fuchsian groups, obtained using hyperbolic geometry, led to generalizations which fueled the growth of combinatorial and geometric group theory in the 20th century.</p>\n\n<p>By work of Dehn and others, cocompact Fuchsian groups are early examples of geometric computations of isoperimetric functions, aka Dehn functions, which Dehn discovered by working with group invariant polygonal tilings of the hyperbolic plane. </p>\n\n<p>Cocompact Fuchsian groups are also interesting early examples of groups with solvable word and conjugacy problem, using ideas related to Dehn's algorithm. </p>\n\n<p>Cocompact Fuchsian group (together with finite rank free groups) are the early examples of hyperbolic groups in the sense of Gromov, proved using that the group with its word metric is quasi-isometric to the hyperbolic plane.</p>\n" }, { "answer_id": 260239, "score": 11, "text": "<p>Fuchsian groups are important for the theory of <a href=\"https://en.wikipedia.org/wiki/Dessin_d%27enfant\" rel=\"nofollow noreferrer\">dessin d'enfants</a> in arithmetic geometry. Wolfart wrote a survey here: <a href=\"http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf\" rel=\"nofollow noreferrer\">http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf</a> .</p>\n\n<p>One typical application is the following: There is some $c&gt;0$, such that for infinitely many $g$ there are $&gt;g^{c\\log g}$ non-isomorphic complex curves $C$ of genus $g$ satisfying $|\\mathrm{Aut}(C)|=84(g-1)$. For the proof you connect the number of different curves to the number of normal subgroups of the $(2,3,7)$-triangle group, and then show that Fuchsian groups are virtually surface groups, and have therefore many normal subgroups.</p>\n" }, { "answer_id": 260372, "score": 7, "text": "<pre><code>Are the Fuchsian groups with fixed points interesting from a geometric perspective?\n</code></pre>\n\n<p>Yes, notice $PSL(2,\\mathbb{Z})$ fixes points. The fundamental domain in this case is isometric to a hyperbolic structure on the disk with two cone points. </p>\n\n<p>More generally, if a Fuchsian group $G$ has finite co-volume in $H^2$, then $H^2/G$ will be a hyperbolic 2-orbifold. For example, all of the orientation preserving subgroups of the hyperbolic triangle groups are Fuchsian groups of this type.</p>\n\n<p>In particular, hyperbolic triangle groups have a unique hyperbolic structure, so if 3-manifold has quotient which contains a (orientable) hyperbolic triangle orbifold then the original 3-manifold contains a totally geodesic surface. </p>\n" }, { "answer_id": 429141, "score": 0, "text": "<p>In Poincaré's words (in translation):</p>\n<blockquote>\n<p>The results so obtained as yet give only a very incomplete solution to\nthe problem I set myself, that is, the integration of linear\ndifferential equations. The equations I have called Fuchsian, and\nwhich can be integrated by means of a simple inversion, are just very\nspecial cases of second-order linear equations.</p>\n</blockquote>\n<p>See page 208 of <em>Uniformization of Riemman Surfaces</em> by Henri Paul de Saint-Gervais.</p>\n" } ]
Question: Why are Fuchsian groups interesting? Answer 1: <p>About the relation to fractals: for Fuchsian groups of the first kind, the limit set has Hausdorff dimension 1, i.e., it is not fractal. </p> <p>However, for all other quasifuchsian groups of the first kind, the limit set has Hausdorff dimension strictly bigger than 1, by a theorem of Rufus Bowen.</p> <p><a href="https://i.stack.imgur.com/WmL2K.gif" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/WmL2K.gif" alt="enter image description here"></a></p> Answer 2: <p>Check out <a href="http://rads.stackoverflow.com/amzn/click/1107564743">Indra's Pearls.</a> (Mumford, Series, Wright).</p> Answer 3: <p>Fuchsian groups occur naturally in JSJ-theory. In 3-manifolds they occur as the base Groups of the Seifert pieces and in geometric group theory they occur as (quotients of) enclosing groups (also called QH-subgroups). In the case of hyperbolic groups these QH-subgroup carry in some sense the outer automorphism group of the whole group.</p> Answer 4: <p>Fuchsian groups, particularly those which are cocompact, form the tips of several big mathematical icebergs. To put this less metaphorically, several discoveries about Fuchsian groups, obtained using hyperbolic geometry, led to generalizations which fueled the growth of combinatorial and geometric group theory in the 20th century.</p> <p>By work of Dehn and others, cocompact Fuchsian groups are early examples of geometric computations of isoperimetric functions, aka Dehn functions, which Dehn discovered by working with group invariant polygonal tilings of the hyperbolic plane. </p> <p>Cocompact Fuchsian groups are also interesting early examples of groups with solvable word and conjugacy problem, using ideas related to Dehn's algorithm. </p> <p>Cocompact Fuchsian group (together with finite rank free groups) are the early examples of hyperbolic groups in the sense of Gromov, proved using that the group with its word metric is quasi-isometric to the hyperbolic plane.</p> Answer 5: <p>Fuchsian groups are important for the theory of <a href="https://en.wikipedia.org/wiki/Dessin_d%27enfant" rel="nofollow noreferrer">dessin d'enfants</a> in arithmetic geometry. Wolfart wrote a survey here: <a href="http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf" rel="nofollow noreferrer">http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf</a> .</p> <p>One typical application is the following: There is some $c&gt;0$, such that for infinitely many $g$ there are $&gt;g^{c\log g}$ non-isomorphic complex curves $C$ of genus $g$ satisfying $|\mathrm{Aut}(C)|=84(g-1)$. For the proof you connect the number of different curves to the number of normal subgroups of the $(2,3,7)$-triangle group, and then show that Fuchsian groups are virtually surface groups, and have therefore many normal subgroups.</p> Answer 6: <pre><code>Are the Fuchsian groups with fixed points interesting from a geometric perspective? </code></pre> <p>Yes, notice $PSL(2,\mathbb{Z})$ fixes points. The fundamental domain in this case is isometric to a hyperbolic structure on the disk with two cone points. </p> <p>More generally, if a Fuchsian group $G$ has finite co-volume in $H^2$, then $H^2/G$ will be a hyperbolic 2-orbifold. For example, all of the orientation preserving subgroups of the hyperbolic triangle groups are Fuchsian groups of this type.</p> <p>In particular, hyperbolic triangle groups have a unique hyperbolic structure, so if 3-manifold has quotient which contains a (orientable) hyperbolic triangle orbifold then the original 3-manifold contains a totally geodesic surface. </p> Answer 7: <p>In Poincaré's words (in translation):</p> <blockquote> <p>The results so obtained as yet give only a very incomplete solution to the problem I set myself, that is, the integration of linear differential equations. The equations I have called Fuchsian, and which can be integrated by means of a simple inversion, are just very special cases of second-order linear equations.</p> </blockquote> <p>See page 208 of <em>Uniformization of Riemman Surfaces</em> by Henri Paul de Saint-Gervais.</p> Answer 8: <p>About the relation to fractals: for Fuchsian groups of the first kind, the limit set has Hausdorff dimension 1, i.e., it is not fractal. </p> <p>However, for all other quasifuchsian groups of the first kind, the limit set has Hausdorff dimension strictly bigger than 1, by a theorem of Rufus Bowen.</p> <p><a href="https://i.stack.imgur.com/WmL2K.gif" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/WmL2K.gif" alt="enter image description here"></a></p> Answer 9: <p>Check out <a href="http://rads.stackoverflow.com/amzn/click/1107564743">Indra's Pearls.</a> (Mumford, Series, Wright).</p> Answer 10: <p>Fuchsian groups occur naturally in JSJ-theory. In 3-manifolds they occur as the base Groups of the Seifert pieces and in geometric group theory they occur as (quotients of) enclosing groups (also called QH-subgroups). In the case of hyperbolic groups these QH-subgroup carry in some sense the outer automorphism group of the whole group.</p> Answer 11: <p>Fuchsian groups, particularly those which are cocompact, form the tips of several big mathematical icebergs. To put this less metaphorically, several discoveries about Fuchsian groups, obtained using hyperbolic geometry, led to generalizations which fueled the growth of combinatorial and geometric group theory in the 20th century.</p> <p>By work of Dehn and others, cocompact Fuchsian groups are early examples of geometric computations of isoperimetric functions, aka Dehn functions, which Dehn discovered by working with group invariant polygonal tilings of the hyperbolic plane. </p> <p>Cocompact Fuchsian groups are also interesting early examples of groups with solvable word and conjugacy problem, using ideas related to Dehn's algorithm. </p> <p>Cocompact Fuchsian group (together with finite rank free groups) are the early examples of hyperbolic groups in the sense of Gromov, proved using that the group with its word metric is quasi-isometric to the hyperbolic plane.</p> Answer 12: <p>Fuchsian groups are important for the theory of <a href="https://en.wikipedia.org/wiki/Dessin_d%27enfant" rel="nofollow noreferrer">dessin d'enfants</a> in arithmetic geometry. Wolfart wrote a survey here: <a href="http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf" rel="nofollow noreferrer">http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf</a> .</p> <p>One typical application is the following: There is some $c&gt;0$, such that for infinitely many $g$ there are $&gt;g^{c\log g}$ non-isomorphic complex curves $C$ of genus $g$ satisfying $|\mathrm{Aut}(C)|=84(g-1)$. For the proof you connect the number of different curves to the number of normal subgroups of the $(2,3,7)$-triangle group, and then show that Fuchsian groups are virtually surface groups, and have therefore many normal subgroups.</p> Answer 13: <pre><code>Are the Fuchsian groups with fixed points interesting from a geometric perspective? </code></pre> <p>Yes, notice $PSL(2,\mathbb{Z})$ fixes points. The fundamental domain in this case is isometric to a hyperbolic structure on the disk with two cone points. </p> <p>More generally, if a Fuchsian group $G$ has finite co-volume in $H^2$, then $H^2/G$ will be a hyperbolic 2-orbifold. For example, all of the orientation preserving subgroups of the hyperbolic triangle groups are Fuchsian groups of this type.</p> <p>In particular, hyperbolic triangle groups have a unique hyperbolic structure, so if 3-manifold has quotient which contains a (orientable) hyperbolic triangle orbifold then the original 3-manifold contains a totally geodesic surface. </p> Answer 14: <p>In Poincaré's words (in translation):</p> <blockquote> <p>The results so obtained as yet give only a very incomplete solution to the problem I set myself, that is, the integration of linear differential equations. The equations I have called Fuchsian, and which can be integrated by means of a simple inversion, are just very special cases of second-order linear equations.</p> </blockquote> <p>See page 208 of <em>Uniformization of Riemman Surfaces</em> by Henri Paul de Saint-Gervais.</p> Answer 15: <p>About the relation to fractals: for Fuchsian groups of the first kind, the limit set has Hausdorff dimension 1, i.e., it is not fractal. </p> <p>However, for all other quasifuchsian groups of the first kind, the limit set has Hausdorff dimension strictly bigger than 1, by a theorem of Rufus Bowen.</p> <p><a href="https://i.stack.imgur.com/WmL2K.gif" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/WmL2K.gif" alt="enter image description here"></a></p> Answer 16: <p>Check out <a href="http://rads.stackoverflow.com/amzn/click/1107564743">Indra's Pearls.</a> (Mumford, Series, Wright).</p> Answer 17: <p>Fuchsian groups occur naturally in JSJ-theory. In 3-manifolds they occur as the base Groups of the Seifert pieces and in geometric group theory they occur as (quotients of) enclosing groups (also called QH-subgroups). In the case of hyperbolic groups these QH-subgroup carry in some sense the outer automorphism group of the whole group.</p> Answer 18: <p>Fuchsian groups, particularly those which are cocompact, form the tips of several big mathematical icebergs. To put this less metaphorically, several discoveries about Fuchsian groups, obtained using hyperbolic geometry, led to generalizations which fueled the growth of combinatorial and geometric group theory in the 20th century.</p> <p>By work of Dehn and others, cocompact Fuchsian groups are early examples of geometric computations of isoperimetric functions, aka Dehn functions, which Dehn discovered by working with group invariant polygonal tilings of the hyperbolic plane. </p> <p>Cocompact Fuchsian groups are also interesting early examples of groups with solvable word and conjugacy problem, using ideas related to Dehn's algorithm. </p> <p>Cocompact Fuchsian group (together with finite rank free groups) are the early examples of hyperbolic groups in the sense of Gromov, proved using that the group with its word metric is quasi-isometric to the hyperbolic plane.</p> Answer 19: <p>Fuchsian groups are important for the theory of <a href="https://en.wikipedia.org/wiki/Dessin_d%27enfant" rel="nofollow noreferrer">dessin d'enfants</a> in arithmetic geometry. Wolfart wrote a survey here: <a href="http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf" rel="nofollow noreferrer">http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf</a> .</p> <p>One typical application is the following: There is some $c&gt;0$, such that for infinitely many $g$ there are $&gt;g^{c\log g}$ non-isomorphic complex curves $C$ of genus $g$ satisfying $|\mathrm{Aut}(C)|=84(g-1)$. For the proof you connect the number of different curves to the number of normal subgroups of the $(2,3,7)$-triangle group, and then show that Fuchsian groups are virtually surface groups, and have therefore many normal subgroups.</p> Answer 20: <pre><code>Are the Fuchsian groups with fixed points interesting from a geometric perspective? </code></pre> <p>Yes, notice $PSL(2,\mathbb{Z})$ fixes points. The fundamental domain in this case is isometric to a hyperbolic structure on the disk with two cone points. </p> <p>More generally, if a Fuchsian group $G$ has finite co-volume in $H^2$, then $H^2/G$ will be a hyperbolic 2-orbifold. For example, all of the orientation preserving subgroups of the hyperbolic triangle groups are Fuchsian groups of this type.</p> <p>In particular, hyperbolic triangle groups have a unique hyperbolic structure, so if 3-manifold has quotient which contains a (orientable) hyperbolic triangle orbifold then the original 3-manifold contains a totally geodesic surface. </p> Answer 21: <p>In Poincaré's words (in translation):</p> <blockquote> <p>The results so obtained as yet give only a very incomplete solution to the problem I set myself, that is, the integration of linear differential equations. The equations I have called Fuchsian, and which can be integrated by means of a simple inversion, are just very special cases of second-order linear equations.</p> </blockquote> <p>See page 208 of <em>Uniformization of Riemman Surfaces</em> by Henri Paul de Saint-Gervais.</p> Answer 22: <p>About the relation to fractals: for Fuchsian groups of the first kind, the limit set has Hausdorff dimension 1, i.e., it is not fractal. </p> <p>However, for all other quasifuchsian groups of the first kind, the limit set has Hausdorff dimension strictly bigger than 1, by a theorem of Rufus Bowen.</p> <p><a href="https://i.stack.imgur.com/WmL2K.gif" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/WmL2K.gif" alt="enter image description here"></a></p> Answer 23: <p>Check out <a href="http://rads.stackoverflow.com/amzn/click/1107564743">Indra's Pearls.</a> (Mumford, Series, Wright).</p> Answer 24: <p>Fuchsian groups occur naturally in JSJ-theory. In 3-manifolds they occur as the base Groups of the Seifert pieces and in geometric group theory they occur as (quotients of) enclosing groups (also called QH-subgroups). In the case of hyperbolic groups these QH-subgroup carry in some sense the outer automorphism group of the whole group.</p> Answer 25: <p>Fuchsian groups, particularly those which are cocompact, form the tips of several big mathematical icebergs. To put this less metaphorically, several discoveries about Fuchsian groups, obtained using hyperbolic geometry, led to generalizations which fueled the growth of combinatorial and geometric group theory in the 20th century.</p> <p>By work of Dehn and others, cocompact Fuchsian groups are early examples of geometric computations of isoperimetric functions, aka Dehn functions, which Dehn discovered by working with group invariant polygonal tilings of the hyperbolic plane. </p> <p>Cocompact Fuchsian groups are also interesting early examples of groups with solvable word and conjugacy problem, using ideas related to Dehn's algorithm. </p> <p>Cocompact Fuchsian group (together with finite rank free groups) are the early examples of hyperbolic groups in the sense of Gromov, proved using that the group with its word metric is quasi-isometric to the hyperbolic plane.</p> Answer 26: <p>Fuchsian groups are important for the theory of <a href="https://en.wikipedia.org/wiki/Dessin_d%27enfant" rel="nofollow noreferrer">dessin d'enfants</a> in arithmetic geometry. Wolfart wrote a survey here: <a href="http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf" rel="nofollow noreferrer">http://www.math.uni-frankfurt.de/~wolfart/Artikel/abc.pdf</a> .</p> <p>One typical application is the following: There is some $c&gt;0$, such that for infinitely many $g$ there are $&gt;g^{c\log g}$ non-isomorphic complex curves $C$ of genus $g$ satisfying $|\mathrm{Aut}(C)|=84(g-1)$. For the proof you connect the number of different curves to the number of normal subgroups of the $(2,3,7)$-triangle group, and then show that Fuchsian groups are virtually surface groups, and have therefore many normal subgroups.</p> Answer 27: <pre><code>Are the Fuchsian groups with fixed points interesting from a geometric perspective? </code></pre> <p>Yes, notice $PSL(2,\mathbb{Z})$ fixes points. The fundamental domain in this case is isometric to a hyperbolic structure on the disk with two cone points. </p> <p>More generally, if a Fuchsian group $G$ has finite co-volume in $H^2$, then $H^2/G$ will be a hyperbolic 2-orbifold. For example, all of the orientation preserving subgroups of the hyperbolic triangle groups are Fuchsian groups of this type.</p> <p>In particular, hyperbolic triangle groups have a unique hyperbolic structure, so if 3-manifold has quotient which contains a (orientable) hyperbolic triangle orbifold then the original 3-manifold contains a totally geodesic surface. </p> Answer 28: <p>In Poincaré's words (in translation):</p> <blockquote> <p>The results so obtained as yet give only a very incomplete solution to the problem I set myself, that is, the integration of linear differential equations. The equations I have called Fuchsian, and which can be integrated by means of a simple inversion, are just very special cases of second-order linear equations.</p> </blockquote> <p>See page 208 of <em>Uniformization of Riemman Surfaces</em> by Henri Paul de Saint-Gervais.</p>
An explicit series representation for the analytic tetration with complex height
259,278
29
https://mathoverflow.net/questions/259278/an-explicit-series-representation-for-the-analytic-tetration-with-complex-height
<p><a href="http://en.wikipedia.org/wiki/Tetration" rel="noreferrer">Tetration</a> is the next hyperoperation after more familiar addition, multiplication and exponentiation. It can be seen as a repeated exponentiation, similar to how exponentiation can be seen as a repeated multiplication, and multiplication — as a repeated addition. Tetration is denoted $^n a$, where $a$ is called the base and $n$ is called the height, and is defined for $n\in\mathbb N\cup\{-1,\,0\}$ by the recurrence $$ {^{-1} a} = 0, \quad {^{n+1} a} = a^{\left({^n a}\right)},\tag1$$ so that $${^0 a}=1, \quad {^1 a} = a, \quad {^2 a} = a^a, \quad {^3 a} = a^{a^a}, \, \dots \quad {^n a} = \underbrace{a^{a^{{.^{.^{.^a}}}}}}_{n\,\text{levels}},\tag2$$ where power towers are evaluated from top to bottom. To simplify matters, we restrict our attention to real bases in the interval $1&lt;a&lt;e^{1/e}$. Under this restriction, the sequence $\{^n a\}$ is strictly increasing and converges to a limit: $$^\infty a=\lim_{n\to\infty} {^n a} = \frac{W(-\ln a)}{-\ln a},\tag3$$ where $W(z)$ denotes the principal branch of the <a href="https://en.wikipedia.org/wiki/Lambert_W_function" rel="noreferrer">Lambert W-function</a>. To simplify our following exposition it is convenient to introduce a new variable $q$ linked to $a$: $$q = \ln({^\infty a}) = {^\infty a} \cdot \ln a = -W(-\ln a).\tag4$$ The bounds on $a$ imply $0&lt;q&lt;1$. The value $q$ is closely related to the asymptotic growth of the tetration. In particular, it appears in remarkable limits: $$ \lim_{n\to\infty} \, \frac{{^{n+k} a}-{^n a}}{{^{n-k} a}-{^n a}} = -q^k, \quad \lim_{n\to\infty} \, \frac{{^{n+k} a}-{^\infty a}}{{^n a}-{^\infty a}} = q^k .\tag5$$ Definitions of addition, multiplication and exponentiation for fractional, negative and complex arguments are well-established. A natural extension of the tetration to fractional and complex heights is a long-standing question. It appears to be much less straightforward problem, mainly because exponentiation, unlike addition and multiplication, lacks both commutativity and associativity, and so tetration does not obey simple laws similar to $a^m \cdot a^n = a^{m+n}$ or $(a^m)^n = a^{m\,n}$. </p> <p>Despite these obstacles, several proposals aiming to extend the definition of the tetration to fractional and complex heights have been put forth. Their common goal is to define a function that agrees with the discrete tetration for all positive integer heights, and satisfies the functional equation $$^{z+1} a = a^{\left({^z a}\right)}\tag6$$ for all $z$ in its domain. These conditions alone are not sufficient to determine a unique function, so additional natural restrictions (such as continuity, smoothness or analyticity) have been proposed, aiming to narrow down the set of candidate functions to a single and, in some sense, "the most natural" extension. The Wikipedia <a href="http://en.wikipedia.org/wiki/Tetration" rel="noreferrer">article</a> on tetration is very cautious and indicates that, apparently, so far there is no consensus about which extension is "the one". Some of the proposals seem to lead to the same function, although apparently no rigorous proof of their equivalence has been published yet. The Citizendum <a href="http://en.citizendium.org/wiki/Tetration" rel="noreferrer">article</a> is written in a much bolder style, and refers to tetration as a unique well-defined function. But, overall, the study of tetration seems to be a quite isolated area of mathematics, with few connections to other topics.</p> <p>A few months ago I made some interesting conjectures related to tetration that I hoped I would prove soon. It appears that there are some connections between tetration and the theory of <a href="http://en.wikipedia.org/wiki/Q-analog" rel="noreferrer">$q$-analogs</a>. Unfortunately, I have not made much progress in proving them since then (although I have learnt a lot about $q$-series), so I finally decided to post them here and ask for your help. First, let's recall some definitions.</p> <p>The <a href="http://en.wikipedia.org/wiki/Q-Pochhammer_symbol" rel="noreferrer">$q$-Pochhammer symbol</a>: $$(p;\,q)_\infty = \prod_{k=0}^\infty(1-p\,q^k)\tag7$$ $$(p;\,q)_z = \frac{(p;\,q)_\infty}{(p\,q^z;\,q)_\infty}, \quad z\in\mathbb C\tag8$$ It easily follows that $$(p;\,q)_n = \prod_{k=0}^{n-1}(1-p\,q^k), \quad n\in\mathbb N\tag9$$</p> <p>The <a href="http://en.wikipedia.org/wiki/Gaussian_binomial_coefficient" rel="noreferrer">$q$-binomial coefficients</a> (also known as the Gaussian binomial coefficients): $${r \brack s}_q = \frac{(q;\,q)_r}{(q;\,q)_s \, (q;\,q)_{r-s}}.\tag{10}$$</p> <p>From now on we assume that $a$ and, consequently, $q$ are fixed. Let us define an analytic function $t(z)$ by the following series:</p> <p>$$ t(z) = \sum_{n=0}^\infty \sum_{k=0}^n (-1)^{n-k} \, q^{\binom {n-k} 2} {z \brack n}_q {n \brack k}_q ({^k a}).\tag{11} $$</p> <p>This formula can be seen as a combination of the direct and reverse $q$-binomial transforms. It can be proved by induction that for $n\in\mathbb N, \, t(n)={^n a}$, so $t(n)$ satisfies the functional equation for tetration $(6)$ at least for positive integer arguments. Indeed, partial sums of $(11)$ can be seen as <a href="http://en.wikipedia.org/wiki/Lagrange_polynomial" rel="noreferrer">Lagrange interpolating polynomials</a> in powers of $q^z$ that exactly reproduce tetration values for progressively larger initial segments of $\mathbb N$. Numeric evidence suggests that adding new points to the interpolating polynomial of this form does not result in erratic oscillations of its arcs between existing points (i.e. <a href="http://en.wikipedia.org/wiki/Runge%27s_phenomenon" rel="noreferrer">Runge's phenomenon</a> does not occur), but rather makes them converge to a monotone function satisfying the functional equation for the tetration $(6)$. So we have:</p> <blockquote> <p><strong>Conjecture 1.</strong> The series $t(z)$ converges at least in the half-plane $\Re(z)&gt;-2$ to an analytic function that satisfies the functional equation for tetration $t(z+1)=a^{t(z)}$ for all arguments in its domain of analyticity.</p> <p><strong>Conjecture 2.</strong> The function $t(z)$ is periodic with a purely imaginary period $\tau = 2\pi n i/\ln q$, and admits an analytic continuation to the half-plane $\Re(z) \le -2$ by repeated backward application of the functional equation $(6)$. There it has a countable set of isolated singularities at $z = m + \tau\,n, \,m,n\in\mathbb Z, \, m \le -2$, and a countable set of branch cuts that can be made along horizontal rays at $z = x + \tau\,n, \, n\in\mathbb Z, \, x\in\mathbb R, \, x &lt; -2$.</p> <p><strong>Conjecture 3.</strong> For real $x&gt;-2$, the derivative $t'(x)$ is a <a href="http://en.wikipedia.org/wiki/Completely_monotone_function" rel="noreferrer">completely monotone function</a>.</p> <p><strong>Conjecture 4.</strong> There is only one analytic function that vanishes at $z=-1$ and satisfies both the functional equation for the tetration $(6)$, and the complete monotonicity condition from the Conjecture 3.</p> </blockquote> <p>Related questions: <a href="https://math.stackexchange.com/questions/2001325/coefficients-of-a-formal-power-series-satisfying-expfz-1-fq-z-q">[1]</a><a href="https://mathoverflow.net/questions/57627/is-there-an-elegant-non-recursive-formula-for-these-coefficients-also-how-ca">[2]</a><a href="https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration">[3]</a>.</p>
[ { "answer_id": 259371, "score": 15, "text": "<p>Not sure if this is an answer but a good suggestion of how to show (1) and (2) (and perhaps an actual solution), and a good start for (3). Consider the Newton series formula for tetration. This series converges for all $\\Re(z) &gt; -2$, and is very similar to your formula</p>\n\n<p>$$^za = \\sum_{n=0}^\\infty \\sum_{k=0}^n (-1)^{n-k} \\binom{z}{n}\\binom{n}{k}(^k a)$$</p>\n\n<p>(The op claims this expression diverges, an alternate expression is</p>\n\n<p>$$\\frac{1}{\\Gamma(1-z)}\\Big{(}\\sum_{n=0}^\\infty (^{n+1}a)\\frac{(-1)^n}{n!(n+1-z)} + \\int_1^\\infty \\big{(}\\sum_{n=0}^\\infty (^{n+1}a) \\frac{(-x)^n}{n!}\\big{)}x^{-z}\\,dx\\Big{)}$$</p>\n\n<p>of which the rest of this discussion still applies.)</p>\n\n<p>Now, since this solution is bounded in the half plane $\\Re(z) &gt; 0$ it does have a uniqueness criterion--quite a good one at that too. It is the only solution to the functional equation to do such. In fact, if a solution $F$ exists where $F(z) = O(e^{\\rho|\\Re(z)| + \\tau|\\Im(z)|})$ for $\\tau &lt; \\pi/2$ then $F$ is still the above solution.</p>\n\n<p>What you have presented is a q-analog of the Newton series I just put above. Sadly I cannot find a rigorous paper detailing these types of series; I don't remember where I have seen them, but I have seen them. This suggests to me, that really you are just interpolating $(^na)$ in the same manner the Newton series does, except you are using q-analogs. It is necessary though that $q$ be fixed to the value you put (explanation below).</p>\n\n<p>Consider the rather beautiful fact that your solution is bounded in the right half plane. </p>\n\n<p>Observe if</p>\n\n<p>$[z]_q = \\frac{1-q^z}{1-q}$, $[z]_q$ is periodic in $z$.</p>\n\n<p>$${z \\brack n}_q = \\frac{[z]_q!}{[n]_q![z-n]_q!} = \\frac{[z]_q[z-1]_q...[z-n+1]_q}{[1]_q[2]_q...[n]_q}$$</p>\n\n<p>And each of these functions have a period, namely $2\\pi i/\\log(q)$..</p>\n\n<p>Therefore your final expansion has an imaginary period, it tends to $-W(\\log(a))/\\log(a)$ as $\\Re(z) \\to \\infty$; thus, it is bounded in the right half plane. </p>\n\n<p>From this it follows your function is the standard tetration function (If it converges).</p>\n\n<p>So $t$ is bounded. The first obvious fact is $t(n) = (^n a)$. Next consider the wondrous power of a little known identity theorem for functions bounded in the right half plane. Namely, if $F(z)$ and $G(z)$ are bounded in the right half plane and $F\\Big{|}_{\\mathbb{N}} = G\\Big{|}_{\\mathbb{N}}$ then $F = G$. This follows because of Ramanujan's master theorem.</p>\n\n<p>I'll explain the identity theorem for a second, and then explain how this applies to your case. If $F$ and $G$ are bounded, then surely</p>\n\n<p>$$f(x) = \\sum_{n=0}^\\infty F(n+1)\\frac{(-x)^n}{n!}$$\n$$g(x) = \\sum_{n=0}^\\infty G(n+1)\\frac{(-x)^n}{n!}$$</p>\n\n<p>are both entire functions, and by Ramanujan both satisfy</p>\n\n<p>$$\\int_0^\\infty f(x)x^{z-1}\\,dx = \\Gamma(z)F(1-z)$$\n$$\\int_0^\\infty g(x)x^{z-1}\\,dx = \\Gamma(z)G(1-z)$$</p>\n\n<p>But $f = g$ and therefore $F = G$.</p>\n\n<p>Look at $t(z+1)$ and $a^{t(z)}$, both of these functions are bounded and agree on the naturals ($t(n+1) = a^{t(n)}$), therefore they are equal everywhere.</p>\n\n<p>Now as to the fact that this function is completely monotone: On the tetration forum Sheldon made a great point. I'm not sure of this fact, but I'm betting it's true. If $0 &lt; \\lambda &lt; 1$ then $f(z) = \\lambda^z$ for $\\Re(z) &gt; 0$ is the only completely monotone function on $\\mathbb{R}^+$ such that $\\lambda f(z) = f(z+1)$. Another function would have to be $f(z+\\theta(z))$ for $\\theta$ a 1-periodic function, but this will most assuredly destroy the completely monotone structure. </p>\n\n<p>What's great about the standard solution (the one that is periodic, the one that is the Schroder iteration, the one that is bounded), is that it can be expanded in an exponential series for $\\Re(z) &gt; 0$. This is just Fourier series, essentially.</p>\n\n<p>$$^za = F(z) = \\sum_{k=0}^\\infty a_k \\lambda^{kz}$$</p>\n\n<p>$$F'(x) \\sim C\\lambda^x$$</p>\n\n<p>Since another solution is just $F(z+\\theta(z))$ for $\\theta$ periodic, we can most likely form a contradiction by using our knowledge of the uniqueness of the exponential function under these conditions. </p>\n\n<p>All in all, I mostly just had to weigh in my two cents, but I think all the conjectures you put are true and that your solution is in fact the 'standard' solution, which is: the periodic one, the bounded one, the Schroder one, and hopefully, the completely monotone one. Hopefully the rough proof I gave of (1) and (2) is good enough for you.</p>\n" }, { "answer_id": 259458, "score": 7, "text": "<p>Here is a proof sketch that the only completely monotone tetration is in fact the standard tetration. I am not certain, as I'm a little fuzzy on Sheldon's details on how to show tetrations derivative is completely monotone. As Sheldon wrote in the comments to my previous answer, the standard tetration (the Schroder iteration) is completely monotone. </p>\n\n<p>He shows it here <a href=\"https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129\">https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129</a>. I think using similar tricks we can show this function is unique under these conditions. By all means my proof is not complete, but I think it's on the right track.</p>\n\n<p>First we start with the wondrous identity that</p>\n\n<p>$$^z a = L + \\sum_{n=1}^\\infty a_n(-1)^n \\lambda^{nz}$$</p>\n\n<p>where $a_n(-1)^n &lt; 0$ for all $n$.</p>\n\n<p>Then if</p>\n\n<p>$$F(z) = \\frac{d}{dz}(^za)$$</p>\n\n<p>it is completely monotone. We wish to show this is the sole function to satisfy this criterion.</p>\n\n<p>This will be our starting point. The next fact to consider is that $\\lambda^{x + \\theta(x)}$ for $\\theta$ 1-periodic is only completely monotone if $\\theta$ is constant. I.e: a completely monotone function multiplied by a periodic function which isn't constant is not completely monotone.</p>\n\n<p>Now assume we have another solution $A(x)$ such that $a^{A(x)} = A(x+1)$, $A(0) = 1$ and $A'(x)$ is completely monotone. It follows with very little work, I'll leave it to the reader, that $A(x) = (^{x+\\theta(x)}a)$ for some 1-periodic function $\\theta$. So that $A'(x) = (1+\\theta'(x))F(x+\\theta(x))$. We also know that $1 + \\theta'(x) \\neq 0$ as this would destroy complete monotonicity. </p>\n\n<p>$A(x)$ can be written as</p>\n\n<p>$$A(x) = L + \\sum_{n=1}^\\infty a_n (-1)^n\\lambda^{nx}\\phi(x)^{n}$$</p>\n\n<p>where $\\phi$ is one periodic. By just looking at this function it is rather clear its derivative can't be completely monotone. The $\\phi$ is obviously going to kill the proper behaviour. To show this though, we'll need to use a good trick.</p>\n\n<p>For the moment, let $\\mu = -\\lambda^x \\phi(x)$, then</p>\n\n<p>$$G(\\mu) = L + \\sum_{n=0}^\\infty a_n \\mu^n$$ </p>\n\n<p>This function is \"fully monotonic,\" which is something that Sheldon showed in his post. The values $(-1)^na_n &lt; 0$ (kind of like the opposite of completely monotone). It is in fact the inverse Schroder function of $l(\\mu) = a^\\mu$. Where $l(G(\\lambda \\mu)) = G(\\mu)$.</p>\n\n<p>Now, $-\\mu = \\lambda^x\\phi(x)$ is not completely monotone (unless $\\phi$ is constant), and therefore composing it with $G(\\mu)$ breaks the monotonic structure. Therefore $\\phi = C$ is constant. Therefore $A(x) = (^{x+C}a)$ but $A(0) = 1$, therefore $C = 0$ or is an imaginary period of $(^z a)$. Either way, $A(x) = (^xa)$.</p>\n\n<p>I think I can say, QED.</p>\n" }, { "answer_id": 259546, "score": 5, "text": "<p>I have written and submitted for publication <a href=\"http://tetration.org/IF.pdf\" rel=\"noreferrer\">The Existence and Uniqueness of the Taylor Series of Iterated Functions</a>. It has cleared an initial peer review with three knowledgeable mathematicians and now is in further peer review for publication. The intent is to have a single authoritative work on extending tetration and all the higher hyperoperators to the complex numbers. The uniqueness and existence theorems in the paper guarantee a unique solution to the problems. So at best, other extensions of tetration must be consistent with the extension provided in the paper. Basing the mathematics on iterated functions provides the generality needed to extend all the higher hyperoperators to complex numbers, not just tetration. </p>\n" }, { "answer_id": 259371, "score": 15, "text": "<p>Not sure if this is an answer but a good suggestion of how to show (1) and (2) (and perhaps an actual solution), and a good start for (3). Consider the Newton series formula for tetration. This series converges for all $\\Re(z) &gt; -2$, and is very similar to your formula</p>\n\n<p>$$^za = \\sum_{n=0}^\\infty \\sum_{k=0}^n (-1)^{n-k} \\binom{z}{n}\\binom{n}{k}(^k a)$$</p>\n\n<p>(The op claims this expression diverges, an alternate expression is</p>\n\n<p>$$\\frac{1}{\\Gamma(1-z)}\\Big{(}\\sum_{n=0}^\\infty (^{n+1}a)\\frac{(-1)^n}{n!(n+1-z)} + \\int_1^\\infty \\big{(}\\sum_{n=0}^\\infty (^{n+1}a) \\frac{(-x)^n}{n!}\\big{)}x^{-z}\\,dx\\Big{)}$$</p>\n\n<p>of which the rest of this discussion still applies.)</p>\n\n<p>Now, since this solution is bounded in the half plane $\\Re(z) &gt; 0$ it does have a uniqueness criterion--quite a good one at that too. It is the only solution to the functional equation to do such. In fact, if a solution $F$ exists where $F(z) = O(e^{\\rho|\\Re(z)| + \\tau|\\Im(z)|})$ for $\\tau &lt; \\pi/2$ then $F$ is still the above solution.</p>\n\n<p>What you have presented is a q-analog of the Newton series I just put above. Sadly I cannot find a rigorous paper detailing these types of series; I don't remember where I have seen them, but I have seen them. This suggests to me, that really you are just interpolating $(^na)$ in the same manner the Newton series does, except you are using q-analogs. It is necessary though that $q$ be fixed to the value you put (explanation below).</p>\n\n<p>Consider the rather beautiful fact that your solution is bounded in the right half plane. </p>\n\n<p>Observe if</p>\n\n<p>$[z]_q = \\frac{1-q^z}{1-q}$, $[z]_q$ is periodic in $z$.</p>\n\n<p>$${z \\brack n}_q = \\frac{[z]_q!}{[n]_q![z-n]_q!} = \\frac{[z]_q[z-1]_q...[z-n+1]_q}{[1]_q[2]_q...[n]_q}$$</p>\n\n<p>And each of these functions have a period, namely $2\\pi i/\\log(q)$..</p>\n\n<p>Therefore your final expansion has an imaginary period, it tends to $-W(\\log(a))/\\log(a)$ as $\\Re(z) \\to \\infty$; thus, it is bounded in the right half plane. </p>\n\n<p>From this it follows your function is the standard tetration function (If it converges).</p>\n\n<p>So $t$ is bounded. The first obvious fact is $t(n) = (^n a)$. Next consider the wondrous power of a little known identity theorem for functions bounded in the right half plane. Namely, if $F(z)$ and $G(z)$ are bounded in the right half plane and $F\\Big{|}_{\\mathbb{N}} = G\\Big{|}_{\\mathbb{N}}$ then $F = G$. This follows because of Ramanujan's master theorem.</p>\n\n<p>I'll explain the identity theorem for a second, and then explain how this applies to your case. If $F$ and $G$ are bounded, then surely</p>\n\n<p>$$f(x) = \\sum_{n=0}^\\infty F(n+1)\\frac{(-x)^n}{n!}$$\n$$g(x) = \\sum_{n=0}^\\infty G(n+1)\\frac{(-x)^n}{n!}$$</p>\n\n<p>are both entire functions, and by Ramanujan both satisfy</p>\n\n<p>$$\\int_0^\\infty f(x)x^{z-1}\\,dx = \\Gamma(z)F(1-z)$$\n$$\\int_0^\\infty g(x)x^{z-1}\\,dx = \\Gamma(z)G(1-z)$$</p>\n\n<p>But $f = g$ and therefore $F = G$.</p>\n\n<p>Look at $t(z+1)$ and $a^{t(z)}$, both of these functions are bounded and agree on the naturals ($t(n+1) = a^{t(n)}$), therefore they are equal everywhere.</p>\n\n<p>Now as to the fact that this function is completely monotone: On the tetration forum Sheldon made a great point. I'm not sure of this fact, but I'm betting it's true. If $0 &lt; \\lambda &lt; 1$ then $f(z) = \\lambda^z$ for $\\Re(z) &gt; 0$ is the only completely monotone function on $\\mathbb{R}^+$ such that $\\lambda f(z) = f(z+1)$. Another function would have to be $f(z+\\theta(z))$ for $\\theta$ a 1-periodic function, but this will most assuredly destroy the completely monotone structure. </p>\n\n<p>What's great about the standard solution (the one that is periodic, the one that is the Schroder iteration, the one that is bounded), is that it can be expanded in an exponential series for $\\Re(z) &gt; 0$. This is just Fourier series, essentially.</p>\n\n<p>$$^za = F(z) = \\sum_{k=0}^\\infty a_k \\lambda^{kz}$$</p>\n\n<p>$$F'(x) \\sim C\\lambda^x$$</p>\n\n<p>Since another solution is just $F(z+\\theta(z))$ for $\\theta$ periodic, we can most likely form a contradiction by using our knowledge of the uniqueness of the exponential function under these conditions. </p>\n\n<p>All in all, I mostly just had to weigh in my two cents, but I think all the conjectures you put are true and that your solution is in fact the 'standard' solution, which is: the periodic one, the bounded one, the Schroder one, and hopefully, the completely monotone one. Hopefully the rough proof I gave of (1) and (2) is good enough for you.</p>\n" }, { "answer_id": 259458, "score": 7, "text": "<p>Here is a proof sketch that the only completely monotone tetration is in fact the standard tetration. I am not certain, as I'm a little fuzzy on Sheldon's details on how to show tetrations derivative is completely monotone. As Sheldon wrote in the comments to my previous answer, the standard tetration (the Schroder iteration) is completely monotone. </p>\n\n<p>He shows it here <a href=\"https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129\">https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129</a>. I think using similar tricks we can show this function is unique under these conditions. By all means my proof is not complete, but I think it's on the right track.</p>\n\n<p>First we start with the wondrous identity that</p>\n\n<p>$$^z a = L + \\sum_{n=1}^\\infty a_n(-1)^n \\lambda^{nz}$$</p>\n\n<p>where $a_n(-1)^n &lt; 0$ for all $n$.</p>\n\n<p>Then if</p>\n\n<p>$$F(z) = \\frac{d}{dz}(^za)$$</p>\n\n<p>it is completely monotone. We wish to show this is the sole function to satisfy this criterion.</p>\n\n<p>This will be our starting point. The next fact to consider is that $\\lambda^{x + \\theta(x)}$ for $\\theta$ 1-periodic is only completely monotone if $\\theta$ is constant. I.e: a completely monotone function multiplied by a periodic function which isn't constant is not completely monotone.</p>\n\n<p>Now assume we have another solution $A(x)$ such that $a^{A(x)} = A(x+1)$, $A(0) = 1$ and $A'(x)$ is completely monotone. It follows with very little work, I'll leave it to the reader, that $A(x) = (^{x+\\theta(x)}a)$ for some 1-periodic function $\\theta$. So that $A'(x) = (1+\\theta'(x))F(x+\\theta(x))$. We also know that $1 + \\theta'(x) \\neq 0$ as this would destroy complete monotonicity. </p>\n\n<p>$A(x)$ can be written as</p>\n\n<p>$$A(x) = L + \\sum_{n=1}^\\infty a_n (-1)^n\\lambda^{nx}\\phi(x)^{n}$$</p>\n\n<p>where $\\phi$ is one periodic. By just looking at this function it is rather clear its derivative can't be completely monotone. The $\\phi$ is obviously going to kill the proper behaviour. To show this though, we'll need to use a good trick.</p>\n\n<p>For the moment, let $\\mu = -\\lambda^x \\phi(x)$, then</p>\n\n<p>$$G(\\mu) = L + \\sum_{n=0}^\\infty a_n \\mu^n$$ </p>\n\n<p>This function is \"fully monotonic,\" which is something that Sheldon showed in his post. The values $(-1)^na_n &lt; 0$ (kind of like the opposite of completely monotone). It is in fact the inverse Schroder function of $l(\\mu) = a^\\mu$. Where $l(G(\\lambda \\mu)) = G(\\mu)$.</p>\n\n<p>Now, $-\\mu = \\lambda^x\\phi(x)$ is not completely monotone (unless $\\phi$ is constant), and therefore composing it with $G(\\mu)$ breaks the monotonic structure. Therefore $\\phi = C$ is constant. Therefore $A(x) = (^{x+C}a)$ but $A(0) = 1$, therefore $C = 0$ or is an imaginary period of $(^z a)$. Either way, $A(x) = (^xa)$.</p>\n\n<p>I think I can say, QED.</p>\n" }, { "answer_id": 259546, "score": 5, "text": "<p>I have written and submitted for publication <a href=\"http://tetration.org/IF.pdf\" rel=\"noreferrer\">The Existence and Uniqueness of the Taylor Series of Iterated Functions</a>. It has cleared an initial peer review with three knowledgeable mathematicians and now is in further peer review for publication. The intent is to have a single authoritative work on extending tetration and all the higher hyperoperators to the complex numbers. The uniqueness and existence theorems in the paper guarantee a unique solution to the problems. So at best, other extensions of tetration must be consistent with the extension provided in the paper. Basing the mathematics on iterated functions provides the generality needed to extend all the higher hyperoperators to complex numbers, not just tetration. </p>\n" }, { "answer_id": 259371, "score": 15, "text": "<p>Not sure if this is an answer but a good suggestion of how to show (1) and (2) (and perhaps an actual solution), and a good start for (3). Consider the Newton series formula for tetration. This series converges for all $\\Re(z) &gt; -2$, and is very similar to your formula</p>\n\n<p>$$^za = \\sum_{n=0}^\\infty \\sum_{k=0}^n (-1)^{n-k} \\binom{z}{n}\\binom{n}{k}(^k a)$$</p>\n\n<p>(The op claims this expression diverges, an alternate expression is</p>\n\n<p>$$\\frac{1}{\\Gamma(1-z)}\\Big{(}\\sum_{n=0}^\\infty (^{n+1}a)\\frac{(-1)^n}{n!(n+1-z)} + \\int_1^\\infty \\big{(}\\sum_{n=0}^\\infty (^{n+1}a) \\frac{(-x)^n}{n!}\\big{)}x^{-z}\\,dx\\Big{)}$$</p>\n\n<p>of which the rest of this discussion still applies.)</p>\n\n<p>Now, since this solution is bounded in the half plane $\\Re(z) &gt; 0$ it does have a uniqueness criterion--quite a good one at that too. It is the only solution to the functional equation to do such. In fact, if a solution $F$ exists where $F(z) = O(e^{\\rho|\\Re(z)| + \\tau|\\Im(z)|})$ for $\\tau &lt; \\pi/2$ then $F$ is still the above solution.</p>\n\n<p>What you have presented is a q-analog of the Newton series I just put above. Sadly I cannot find a rigorous paper detailing these types of series; I don't remember where I have seen them, but I have seen them. This suggests to me, that really you are just interpolating $(^na)$ in the same manner the Newton series does, except you are using q-analogs. It is necessary though that $q$ be fixed to the value you put (explanation below).</p>\n\n<p>Consider the rather beautiful fact that your solution is bounded in the right half plane. </p>\n\n<p>Observe if</p>\n\n<p>$[z]_q = \\frac{1-q^z}{1-q}$, $[z]_q$ is periodic in $z$.</p>\n\n<p>$${z \\brack n}_q = \\frac{[z]_q!}{[n]_q![z-n]_q!} = \\frac{[z]_q[z-1]_q...[z-n+1]_q}{[1]_q[2]_q...[n]_q}$$</p>\n\n<p>And each of these functions have a period, namely $2\\pi i/\\log(q)$..</p>\n\n<p>Therefore your final expansion has an imaginary period, it tends to $-W(\\log(a))/\\log(a)$ as $\\Re(z) \\to \\infty$; thus, it is bounded in the right half plane. </p>\n\n<p>From this it follows your function is the standard tetration function (If it converges).</p>\n\n<p>So $t$ is bounded. The first obvious fact is $t(n) = (^n a)$. Next consider the wondrous power of a little known identity theorem for functions bounded in the right half plane. Namely, if $F(z)$ and $G(z)$ are bounded in the right half plane and $F\\Big{|}_{\\mathbb{N}} = G\\Big{|}_{\\mathbb{N}}$ then $F = G$. This follows because of Ramanujan's master theorem.</p>\n\n<p>I'll explain the identity theorem for a second, and then explain how this applies to your case. If $F$ and $G$ are bounded, then surely</p>\n\n<p>$$f(x) = \\sum_{n=0}^\\infty F(n+1)\\frac{(-x)^n}{n!}$$\n$$g(x) = \\sum_{n=0}^\\infty G(n+1)\\frac{(-x)^n}{n!}$$</p>\n\n<p>are both entire functions, and by Ramanujan both satisfy</p>\n\n<p>$$\\int_0^\\infty f(x)x^{z-1}\\,dx = \\Gamma(z)F(1-z)$$\n$$\\int_0^\\infty g(x)x^{z-1}\\,dx = \\Gamma(z)G(1-z)$$</p>\n\n<p>But $f = g$ and therefore $F = G$.</p>\n\n<p>Look at $t(z+1)$ and $a^{t(z)}$, both of these functions are bounded and agree on the naturals ($t(n+1) = a^{t(n)}$), therefore they are equal everywhere.</p>\n\n<p>Now as to the fact that this function is completely monotone: On the tetration forum Sheldon made a great point. I'm not sure of this fact, but I'm betting it's true. If $0 &lt; \\lambda &lt; 1$ then $f(z) = \\lambda^z$ for $\\Re(z) &gt; 0$ is the only completely monotone function on $\\mathbb{R}^+$ such that $\\lambda f(z) = f(z+1)$. Another function would have to be $f(z+\\theta(z))$ for $\\theta$ a 1-periodic function, but this will most assuredly destroy the completely monotone structure. </p>\n\n<p>What's great about the standard solution (the one that is periodic, the one that is the Schroder iteration, the one that is bounded), is that it can be expanded in an exponential series for $\\Re(z) &gt; 0$. This is just Fourier series, essentially.</p>\n\n<p>$$^za = F(z) = \\sum_{k=0}^\\infty a_k \\lambda^{kz}$$</p>\n\n<p>$$F'(x) \\sim C\\lambda^x$$</p>\n\n<p>Since another solution is just $F(z+\\theta(z))$ for $\\theta$ periodic, we can most likely form a contradiction by using our knowledge of the uniqueness of the exponential function under these conditions. </p>\n\n<p>All in all, I mostly just had to weigh in my two cents, but I think all the conjectures you put are true and that your solution is in fact the 'standard' solution, which is: the periodic one, the bounded one, the Schroder one, and hopefully, the completely monotone one. Hopefully the rough proof I gave of (1) and (2) is good enough for you.</p>\n" }, { "answer_id": 259458, "score": 7, "text": "<p>Here is a proof sketch that the only completely monotone tetration is in fact the standard tetration. I am not certain, as I'm a little fuzzy on Sheldon's details on how to show tetrations derivative is completely monotone. As Sheldon wrote in the comments to my previous answer, the standard tetration (the Schroder iteration) is completely monotone. </p>\n\n<p>He shows it here <a href=\"https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129\">https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129</a>. I think using similar tricks we can show this function is unique under these conditions. By all means my proof is not complete, but I think it's on the right track.</p>\n\n<p>First we start with the wondrous identity that</p>\n\n<p>$$^z a = L + \\sum_{n=1}^\\infty a_n(-1)^n \\lambda^{nz}$$</p>\n\n<p>where $a_n(-1)^n &lt; 0$ for all $n$.</p>\n\n<p>Then if</p>\n\n<p>$$F(z) = \\frac{d}{dz}(^za)$$</p>\n\n<p>it is completely monotone. We wish to show this is the sole function to satisfy this criterion.</p>\n\n<p>This will be our starting point. The next fact to consider is that $\\lambda^{x + \\theta(x)}$ for $\\theta$ 1-periodic is only completely monotone if $\\theta$ is constant. I.e: a completely monotone function multiplied by a periodic function which isn't constant is not completely monotone.</p>\n\n<p>Now assume we have another solution $A(x)$ such that $a^{A(x)} = A(x+1)$, $A(0) = 1$ and $A'(x)$ is completely monotone. It follows with very little work, I'll leave it to the reader, that $A(x) = (^{x+\\theta(x)}a)$ for some 1-periodic function $\\theta$. So that $A'(x) = (1+\\theta'(x))F(x+\\theta(x))$. We also know that $1 + \\theta'(x) \\neq 0$ as this would destroy complete monotonicity. </p>\n\n<p>$A(x)$ can be written as</p>\n\n<p>$$A(x) = L + \\sum_{n=1}^\\infty a_n (-1)^n\\lambda^{nx}\\phi(x)^{n}$$</p>\n\n<p>where $\\phi$ is one periodic. By just looking at this function it is rather clear its derivative can't be completely monotone. The $\\phi$ is obviously going to kill the proper behaviour. To show this though, we'll need to use a good trick.</p>\n\n<p>For the moment, let $\\mu = -\\lambda^x \\phi(x)$, then</p>\n\n<p>$$G(\\mu) = L + \\sum_{n=0}^\\infty a_n \\mu^n$$ </p>\n\n<p>This function is \"fully monotonic,\" which is something that Sheldon showed in his post. The values $(-1)^na_n &lt; 0$ (kind of like the opposite of completely monotone). It is in fact the inverse Schroder function of $l(\\mu) = a^\\mu$. Where $l(G(\\lambda \\mu)) = G(\\mu)$.</p>\n\n<p>Now, $-\\mu = \\lambda^x\\phi(x)$ is not completely monotone (unless $\\phi$ is constant), and therefore composing it with $G(\\mu)$ breaks the monotonic structure. Therefore $\\phi = C$ is constant. Therefore $A(x) = (^{x+C}a)$ but $A(0) = 1$, therefore $C = 0$ or is an imaginary period of $(^z a)$. Either way, $A(x) = (^xa)$.</p>\n\n<p>I think I can say, QED.</p>\n" }, { "answer_id": 259546, "score": 5, "text": "<p>I have written and submitted for publication <a href=\"http://tetration.org/IF.pdf\" rel=\"noreferrer\">The Existence and Uniqueness of the Taylor Series of Iterated Functions</a>. It has cleared an initial peer review with three knowledgeable mathematicians and now is in further peer review for publication. The intent is to have a single authoritative work on extending tetration and all the higher hyperoperators to the complex numbers. The uniqueness and existence theorems in the paper guarantee a unique solution to the problems. So at best, other extensions of tetration must be consistent with the extension provided in the paper. Basing the mathematics on iterated functions provides the generality needed to extend all the higher hyperoperators to complex numbers, not just tetration. </p>\n" }, { "answer_id": 259371, "score": 15, "text": "<p>Not sure if this is an answer but a good suggestion of how to show (1) and (2) (and perhaps an actual solution), and a good start for (3). Consider the Newton series formula for tetration. This series converges for all $\\Re(z) &gt; -2$, and is very similar to your formula</p>\n\n<p>$$^za = \\sum_{n=0}^\\infty \\sum_{k=0}^n (-1)^{n-k} \\binom{z}{n}\\binom{n}{k}(^k a)$$</p>\n\n<p>(The op claims this expression diverges, an alternate expression is</p>\n\n<p>$$\\frac{1}{\\Gamma(1-z)}\\Big{(}\\sum_{n=0}^\\infty (^{n+1}a)\\frac{(-1)^n}{n!(n+1-z)} + \\int_1^\\infty \\big{(}\\sum_{n=0}^\\infty (^{n+1}a) \\frac{(-x)^n}{n!}\\big{)}x^{-z}\\,dx\\Big{)}$$</p>\n\n<p>of which the rest of this discussion still applies.)</p>\n\n<p>Now, since this solution is bounded in the half plane $\\Re(z) &gt; 0$ it does have a uniqueness criterion--quite a good one at that too. It is the only solution to the functional equation to do such. In fact, if a solution $F$ exists where $F(z) = O(e^{\\rho|\\Re(z)| + \\tau|\\Im(z)|})$ for $\\tau &lt; \\pi/2$ then $F$ is still the above solution.</p>\n\n<p>What you have presented is a q-analog of the Newton series I just put above. Sadly I cannot find a rigorous paper detailing these types of series; I don't remember where I have seen them, but I have seen them. This suggests to me, that really you are just interpolating $(^na)$ in the same manner the Newton series does, except you are using q-analogs. It is necessary though that $q$ be fixed to the value you put (explanation below).</p>\n\n<p>Consider the rather beautiful fact that your solution is bounded in the right half plane. </p>\n\n<p>Observe if</p>\n\n<p>$[z]_q = \\frac{1-q^z}{1-q}$, $[z]_q$ is periodic in $z$.</p>\n\n<p>$${z \\brack n}_q = \\frac{[z]_q!}{[n]_q![z-n]_q!} = \\frac{[z]_q[z-1]_q...[z-n+1]_q}{[1]_q[2]_q...[n]_q}$$</p>\n\n<p>And each of these functions have a period, namely $2\\pi i/\\log(q)$..</p>\n\n<p>Therefore your final expansion has an imaginary period, it tends to $-W(\\log(a))/\\log(a)$ as $\\Re(z) \\to \\infty$; thus, it is bounded in the right half plane. </p>\n\n<p>From this it follows your function is the standard tetration function (If it converges).</p>\n\n<p>So $t$ is bounded. The first obvious fact is $t(n) = (^n a)$. Next consider the wondrous power of a little known identity theorem for functions bounded in the right half plane. Namely, if $F(z)$ and $G(z)$ are bounded in the right half plane and $F\\Big{|}_{\\mathbb{N}} = G\\Big{|}_{\\mathbb{N}}$ then $F = G$. This follows because of Ramanujan's master theorem.</p>\n\n<p>I'll explain the identity theorem for a second, and then explain how this applies to your case. If $F$ and $G$ are bounded, then surely</p>\n\n<p>$$f(x) = \\sum_{n=0}^\\infty F(n+1)\\frac{(-x)^n}{n!}$$\n$$g(x) = \\sum_{n=0}^\\infty G(n+1)\\frac{(-x)^n}{n!}$$</p>\n\n<p>are both entire functions, and by Ramanujan both satisfy</p>\n\n<p>$$\\int_0^\\infty f(x)x^{z-1}\\,dx = \\Gamma(z)F(1-z)$$\n$$\\int_0^\\infty g(x)x^{z-1}\\,dx = \\Gamma(z)G(1-z)$$</p>\n\n<p>But $f = g$ and therefore $F = G$.</p>\n\n<p>Look at $t(z+1)$ and $a^{t(z)}$, both of these functions are bounded and agree on the naturals ($t(n+1) = a^{t(n)}$), therefore they are equal everywhere.</p>\n\n<p>Now as to the fact that this function is completely monotone: On the tetration forum Sheldon made a great point. I'm not sure of this fact, but I'm betting it's true. If $0 &lt; \\lambda &lt; 1$ then $f(z) = \\lambda^z$ for $\\Re(z) &gt; 0$ is the only completely monotone function on $\\mathbb{R}^+$ such that $\\lambda f(z) = f(z+1)$. Another function would have to be $f(z+\\theta(z))$ for $\\theta$ a 1-periodic function, but this will most assuredly destroy the completely monotone structure. </p>\n\n<p>What's great about the standard solution (the one that is periodic, the one that is the Schroder iteration, the one that is bounded), is that it can be expanded in an exponential series for $\\Re(z) &gt; 0$. This is just Fourier series, essentially.</p>\n\n<p>$$^za = F(z) = \\sum_{k=0}^\\infty a_k \\lambda^{kz}$$</p>\n\n<p>$$F'(x) \\sim C\\lambda^x$$</p>\n\n<p>Since another solution is just $F(z+\\theta(z))$ for $\\theta$ periodic, we can most likely form a contradiction by using our knowledge of the uniqueness of the exponential function under these conditions. </p>\n\n<p>All in all, I mostly just had to weigh in my two cents, but I think all the conjectures you put are true and that your solution is in fact the 'standard' solution, which is: the periodic one, the bounded one, the Schroder one, and hopefully, the completely monotone one. Hopefully the rough proof I gave of (1) and (2) is good enough for you.</p>\n" }, { "answer_id": 259458, "score": 7, "text": "<p>Here is a proof sketch that the only completely monotone tetration is in fact the standard tetration. I am not certain, as I'm a little fuzzy on Sheldon's details on how to show tetrations derivative is completely monotone. As Sheldon wrote in the comments to my previous answer, the standard tetration (the Schroder iteration) is completely monotone. </p>\n\n<p>He shows it here <a href=\"https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129\">https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129</a>. I think using similar tricks we can show this function is unique under these conditions. By all means my proof is not complete, but I think it's on the right track.</p>\n\n<p>First we start with the wondrous identity that</p>\n\n<p>$$^z a = L + \\sum_{n=1}^\\infty a_n(-1)^n \\lambda^{nz}$$</p>\n\n<p>where $a_n(-1)^n &lt; 0$ for all $n$.</p>\n\n<p>Then if</p>\n\n<p>$$F(z) = \\frac{d}{dz}(^za)$$</p>\n\n<p>it is completely monotone. We wish to show this is the sole function to satisfy this criterion.</p>\n\n<p>This will be our starting point. The next fact to consider is that $\\lambda^{x + \\theta(x)}$ for $\\theta$ 1-periodic is only completely monotone if $\\theta$ is constant. I.e: a completely monotone function multiplied by a periodic function which isn't constant is not completely monotone.</p>\n\n<p>Now assume we have another solution $A(x)$ such that $a^{A(x)} = A(x+1)$, $A(0) = 1$ and $A'(x)$ is completely monotone. It follows with very little work, I'll leave it to the reader, that $A(x) = (^{x+\\theta(x)}a)$ for some 1-periodic function $\\theta$. So that $A'(x) = (1+\\theta'(x))F(x+\\theta(x))$. We also know that $1 + \\theta'(x) \\neq 0$ as this would destroy complete monotonicity. </p>\n\n<p>$A(x)$ can be written as</p>\n\n<p>$$A(x) = L + \\sum_{n=1}^\\infty a_n (-1)^n\\lambda^{nx}\\phi(x)^{n}$$</p>\n\n<p>where $\\phi$ is one periodic. By just looking at this function it is rather clear its derivative can't be completely monotone. The $\\phi$ is obviously going to kill the proper behaviour. To show this though, we'll need to use a good trick.</p>\n\n<p>For the moment, let $\\mu = -\\lambda^x \\phi(x)$, then</p>\n\n<p>$$G(\\mu) = L + \\sum_{n=0}^\\infty a_n \\mu^n$$ </p>\n\n<p>This function is \"fully monotonic,\" which is something that Sheldon showed in his post. The values $(-1)^na_n &lt; 0$ (kind of like the opposite of completely monotone). It is in fact the inverse Schroder function of $l(\\mu) = a^\\mu$. Where $l(G(\\lambda \\mu)) = G(\\mu)$.</p>\n\n<p>Now, $-\\mu = \\lambda^x\\phi(x)$ is not completely monotone (unless $\\phi$ is constant), and therefore composing it with $G(\\mu)$ breaks the monotonic structure. Therefore $\\phi = C$ is constant. Therefore $A(x) = (^{x+C}a)$ but $A(0) = 1$, therefore $C = 0$ or is an imaginary period of $(^z a)$. Either way, $A(x) = (^xa)$.</p>\n\n<p>I think I can say, QED.</p>\n" }, { "answer_id": 259546, "score": 5, "text": "<p>I have written and submitted for publication <a href=\"http://tetration.org/IF.pdf\" rel=\"noreferrer\">The Existence and Uniqueness of the Taylor Series of Iterated Functions</a>. It has cleared an initial peer review with three knowledgeable mathematicians and now is in further peer review for publication. The intent is to have a single authoritative work on extending tetration and all the higher hyperoperators to the complex numbers. The uniqueness and existence theorems in the paper guarantee a unique solution to the problems. So at best, other extensions of tetration must be consistent with the extension provided in the paper. Basing the mathematics on iterated functions provides the generality needed to extend all the higher hyperoperators to complex numbers, not just tetration. </p>\n" } ]
Question: An explicit series representation for the analytic tetration with complex height Answer 1: <p>Not sure if this is an answer but a good suggestion of how to show (1) and (2) (and perhaps an actual solution), and a good start for (3). Consider the Newton series formula for tetration. This series converges for all $\Re(z) &gt; -2$, and is very similar to your formula</p> <p>$$^za = \sum_{n=0}^\infty \sum_{k=0}^n (-1)^{n-k} \binom{z}{n}\binom{n}{k}(^k a)$$</p> <p>(The op claims this expression diverges, an alternate expression is</p> <p>$$\frac{1}{\Gamma(1-z)}\Big{(}\sum_{n=0}^\infty (^{n+1}a)\frac{(-1)^n}{n!(n+1-z)} + \int_1^\infty \big{(}\sum_{n=0}^\infty (^{n+1}a) \frac{(-x)^n}{n!}\big{)}x^{-z}\,dx\Big{)}$$</p> <p>of which the rest of this discussion still applies.)</p> <p>Now, since this solution is bounded in the half plane $\Re(z) &gt; 0$ it does have a uniqueness criterion--quite a good one at that too. It is the only solution to the functional equation to do such. In fact, if a solution $F$ exists where $F(z) = O(e^{\rho|\Re(z)| + \tau|\Im(z)|})$ for $\tau &lt; \pi/2$ then $F$ is still the above solution.</p> <p>What you have presented is a q-analog of the Newton series I just put above. Sadly I cannot find a rigorous paper detailing these types of series; I don't remember where I have seen them, but I have seen them. This suggests to me, that really you are just interpolating $(^na)$ in the same manner the Newton series does, except you are using q-analogs. It is necessary though that $q$ be fixed to the value you put (explanation below).</p> <p>Consider the rather beautiful fact that your solution is bounded in the right half plane. </p> <p>Observe if</p> <p>$[z]_q = \frac{1-q^z}{1-q}$, $[z]_q$ is periodic in $z$.</p> <p>$${z \brack n}_q = \frac{[z]_q!}{[n]_q![z-n]_q!} = \frac{[z]_q[z-1]_q...[z-n+1]_q}{[1]_q[2]_q...[n]_q}$$</p> <p>And each of these functions have a period, namely $2\pi i/\log(q)$..</p> <p>Therefore your final expansion has an imaginary period, it tends to $-W(\log(a))/\log(a)$ as $\Re(z) \to \infty$; thus, it is bounded in the right half plane. </p> <p>From this it follows your function is the standard tetration function (If it converges).</p> <p>So $t$ is bounded. The first obvious fact is $t(n) = (^n a)$. Next consider the wondrous power of a little known identity theorem for functions bounded in the right half plane. Namely, if $F(z)$ and $G(z)$ are bounded in the right half plane and $F\Big{|}_{\mathbb{N}} = G\Big{|}_{\mathbb{N}}$ then $F = G$. This follows because of Ramanujan's master theorem.</p> <p>I'll explain the identity theorem for a second, and then explain how this applies to your case. If $F$ and $G$ are bounded, then surely</p> <p>$$f(x) = \sum_{n=0}^\infty F(n+1)\frac{(-x)^n}{n!}$$ $$g(x) = \sum_{n=0}^\infty G(n+1)\frac{(-x)^n}{n!}$$</p> <p>are both entire functions, and by Ramanujan both satisfy</p> <p>$$\int_0^\infty f(x)x^{z-1}\,dx = \Gamma(z)F(1-z)$$ $$\int_0^\infty g(x)x^{z-1}\,dx = \Gamma(z)G(1-z)$$</p> <p>But $f = g$ and therefore $F = G$.</p> <p>Look at $t(z+1)$ and $a^{t(z)}$, both of these functions are bounded and agree on the naturals ($t(n+1) = a^{t(n)}$), therefore they are equal everywhere.</p> <p>Now as to the fact that this function is completely monotone: On the tetration forum Sheldon made a great point. I'm not sure of this fact, but I'm betting it's true. If $0 &lt; \lambda &lt; 1$ then $f(z) = \lambda^z$ for $\Re(z) &gt; 0$ is the only completely monotone function on $\mathbb{R}^+$ such that $\lambda f(z) = f(z+1)$. Another function would have to be $f(z+\theta(z))$ for $\theta$ a 1-periodic function, but this will most assuredly destroy the completely monotone structure. </p> <p>What's great about the standard solution (the one that is periodic, the one that is the Schroder iteration, the one that is bounded), is that it can be expanded in an exponential series for $\Re(z) &gt; 0$. This is just Fourier series, essentially.</p> <p>$$^za = F(z) = \sum_{k=0}^\infty a_k \lambda^{kz}$$</p> <p>$$F'(x) \sim C\lambda^x$$</p> <p>Since another solution is just $F(z+\theta(z))$ for $\theta$ periodic, we can most likely form a contradiction by using our knowledge of the uniqueness of the exponential function under these conditions. </p> <p>All in all, I mostly just had to weigh in my two cents, but I think all the conjectures you put are true and that your solution is in fact the 'standard' solution, which is: the periodic one, the bounded one, the Schroder one, and hopefully, the completely monotone one. Hopefully the rough proof I gave of (1) and (2) is good enough for you.</p> Answer 2: <p>Here is a proof sketch that the only completely monotone tetration is in fact the standard tetration. I am not certain, as I'm a little fuzzy on Sheldon's details on how to show tetrations derivative is completely monotone. As Sheldon wrote in the comments to my previous answer, the standard tetration (the Schroder iteration) is completely monotone. </p> <p>He shows it here <a href="https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129">https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129</a>. I think using similar tricks we can show this function is unique under these conditions. By all means my proof is not complete, but I think it's on the right track.</p> <p>First we start with the wondrous identity that</p> <p>$$^z a = L + \sum_{n=1}^\infty a_n(-1)^n \lambda^{nz}$$</p> <p>where $a_n(-1)^n &lt; 0$ for all $n$.</p> <p>Then if</p> <p>$$F(z) = \frac{d}{dz}(^za)$$</p> <p>it is completely monotone. We wish to show this is the sole function to satisfy this criterion.</p> <p>This will be our starting point. The next fact to consider is that $\lambda^{x + \theta(x)}$ for $\theta$ 1-periodic is only completely monotone if $\theta$ is constant. I.e: a completely monotone function multiplied by a periodic function which isn't constant is not completely monotone.</p> <p>Now assume we have another solution $A(x)$ such that $a^{A(x)} = A(x+1)$, $A(0) = 1$ and $A'(x)$ is completely monotone. It follows with very little work, I'll leave it to the reader, that $A(x) = (^{x+\theta(x)}a)$ for some 1-periodic function $\theta$. So that $A'(x) = (1+\theta'(x))F(x+\theta(x))$. We also know that $1 + \theta'(x) \neq 0$ as this would destroy complete monotonicity. </p> <p>$A(x)$ can be written as</p> <p>$$A(x) = L + \sum_{n=1}^\infty a_n (-1)^n\lambda^{nx}\phi(x)^{n}$$</p> <p>where $\phi$ is one periodic. By just looking at this function it is rather clear its derivative can't be completely monotone. The $\phi$ is obviously going to kill the proper behaviour. To show this though, we'll need to use a good trick.</p> <p>For the moment, let $\mu = -\lambda^x \phi(x)$, then</p> <p>$$G(\mu) = L + \sum_{n=0}^\infty a_n \mu^n$$ </p> <p>This function is "fully monotonic," which is something that Sheldon showed in his post. The values $(-1)^na_n &lt; 0$ (kind of like the opposite of completely monotone). It is in fact the inverse Schroder function of $l(\mu) = a^\mu$. Where $l(G(\lambda \mu)) = G(\mu)$.</p> <p>Now, $-\mu = \lambda^x\phi(x)$ is not completely monotone (unless $\phi$ is constant), and therefore composing it with $G(\mu)$ breaks the monotonic structure. Therefore $\phi = C$ is constant. Therefore $A(x) = (^{x+C}a)$ but $A(0) = 1$, therefore $C = 0$ or is an imaginary period of $(^z a)$. Either way, $A(x) = (^xa)$.</p> <p>I think I can say, QED.</p> Answer 3: <p>I have written and submitted for publication <a href="http://tetration.org/IF.pdf" rel="noreferrer">The Existence and Uniqueness of the Taylor Series of Iterated Functions</a>. It has cleared an initial peer review with three knowledgeable mathematicians and now is in further peer review for publication. The intent is to have a single authoritative work on extending tetration and all the higher hyperoperators to the complex numbers. The uniqueness and existence theorems in the paper guarantee a unique solution to the problems. So at best, other extensions of tetration must be consistent with the extension provided in the paper. Basing the mathematics on iterated functions provides the generality needed to extend all the higher hyperoperators to complex numbers, not just tetration. </p> Answer 4: <p>Not sure if this is an answer but a good suggestion of how to show (1) and (2) (and perhaps an actual solution), and a good start for (3). Consider the Newton series formula for tetration. This series converges for all $\Re(z) &gt; -2$, and is very similar to your formula</p> <p>$$^za = \sum_{n=0}^\infty \sum_{k=0}^n (-1)^{n-k} \binom{z}{n}\binom{n}{k}(^k a)$$</p> <p>(The op claims this expression diverges, an alternate expression is</p> <p>$$\frac{1}{\Gamma(1-z)}\Big{(}\sum_{n=0}^\infty (^{n+1}a)\frac{(-1)^n}{n!(n+1-z)} + \int_1^\infty \big{(}\sum_{n=0}^\infty (^{n+1}a) \frac{(-x)^n}{n!}\big{)}x^{-z}\,dx\Big{)}$$</p> <p>of which the rest of this discussion still applies.)</p> <p>Now, since this solution is bounded in the half plane $\Re(z) &gt; 0$ it does have a uniqueness criterion--quite a good one at that too. It is the only solution to the functional equation to do such. In fact, if a solution $F$ exists where $F(z) = O(e^{\rho|\Re(z)| + \tau|\Im(z)|})$ for $\tau &lt; \pi/2$ then $F$ is still the above solution.</p> <p>What you have presented is a q-analog of the Newton series I just put above. Sadly I cannot find a rigorous paper detailing these types of series; I don't remember where I have seen them, but I have seen them. This suggests to me, that really you are just interpolating $(^na)$ in the same manner the Newton series does, except you are using q-analogs. It is necessary though that $q$ be fixed to the value you put (explanation below).</p> <p>Consider the rather beautiful fact that your solution is bounded in the right half plane. </p> <p>Observe if</p> <p>$[z]_q = \frac{1-q^z}{1-q}$, $[z]_q$ is periodic in $z$.</p> <p>$${z \brack n}_q = \frac{[z]_q!}{[n]_q![z-n]_q!} = \frac{[z]_q[z-1]_q...[z-n+1]_q}{[1]_q[2]_q...[n]_q}$$</p> <p>And each of these functions have a period, namely $2\pi i/\log(q)$..</p> <p>Therefore your final expansion has an imaginary period, it tends to $-W(\log(a))/\log(a)$ as $\Re(z) \to \infty$; thus, it is bounded in the right half plane. </p> <p>From this it follows your function is the standard tetration function (If it converges).</p> <p>So $t$ is bounded. The first obvious fact is $t(n) = (^n a)$. Next consider the wondrous power of a little known identity theorem for functions bounded in the right half plane. Namely, if $F(z)$ and $G(z)$ are bounded in the right half plane and $F\Big{|}_{\mathbb{N}} = G\Big{|}_{\mathbb{N}}$ then $F = G$. This follows because of Ramanujan's master theorem.</p> <p>I'll explain the identity theorem for a second, and then explain how this applies to your case. If $F$ and $G$ are bounded, then surely</p> <p>$$f(x) = \sum_{n=0}^\infty F(n+1)\frac{(-x)^n}{n!}$$ $$g(x) = \sum_{n=0}^\infty G(n+1)\frac{(-x)^n}{n!}$$</p> <p>are both entire functions, and by Ramanujan both satisfy</p> <p>$$\int_0^\infty f(x)x^{z-1}\,dx = \Gamma(z)F(1-z)$$ $$\int_0^\infty g(x)x^{z-1}\,dx = \Gamma(z)G(1-z)$$</p> <p>But $f = g$ and therefore $F = G$.</p> <p>Look at $t(z+1)$ and $a^{t(z)}$, both of these functions are bounded and agree on the naturals ($t(n+1) = a^{t(n)}$), therefore they are equal everywhere.</p> <p>Now as to the fact that this function is completely monotone: On the tetration forum Sheldon made a great point. I'm not sure of this fact, but I'm betting it's true. If $0 &lt; \lambda &lt; 1$ then $f(z) = \lambda^z$ for $\Re(z) &gt; 0$ is the only completely monotone function on $\mathbb{R}^+$ such that $\lambda f(z) = f(z+1)$. Another function would have to be $f(z+\theta(z))$ for $\theta$ a 1-periodic function, but this will most assuredly destroy the completely monotone structure. </p> <p>What's great about the standard solution (the one that is periodic, the one that is the Schroder iteration, the one that is bounded), is that it can be expanded in an exponential series for $\Re(z) &gt; 0$. This is just Fourier series, essentially.</p> <p>$$^za = F(z) = \sum_{k=0}^\infty a_k \lambda^{kz}$$</p> <p>$$F'(x) \sim C\lambda^x$$</p> <p>Since another solution is just $F(z+\theta(z))$ for $\theta$ periodic, we can most likely form a contradiction by using our knowledge of the uniqueness of the exponential function under these conditions. </p> <p>All in all, I mostly just had to weigh in my two cents, but I think all the conjectures you put are true and that your solution is in fact the 'standard' solution, which is: the periodic one, the bounded one, the Schroder one, and hopefully, the completely monotone one. Hopefully the rough proof I gave of (1) and (2) is good enough for you.</p> Answer 5: <p>Here is a proof sketch that the only completely monotone tetration is in fact the standard tetration. I am not certain, as I'm a little fuzzy on Sheldon's details on how to show tetrations derivative is completely monotone. As Sheldon wrote in the comments to my previous answer, the standard tetration (the Schroder iteration) is completely monotone. </p> <p>He shows it here <a href="https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129">https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129</a>. I think using similar tricks we can show this function is unique under these conditions. By all means my proof is not complete, but I think it's on the right track.</p> <p>First we start with the wondrous identity that</p> <p>$$^z a = L + \sum_{n=1}^\infty a_n(-1)^n \lambda^{nz}$$</p> <p>where $a_n(-1)^n &lt; 0$ for all $n$.</p> <p>Then if</p> <p>$$F(z) = \frac{d}{dz}(^za)$$</p> <p>it is completely monotone. We wish to show this is the sole function to satisfy this criterion.</p> <p>This will be our starting point. The next fact to consider is that $\lambda^{x + \theta(x)}$ for $\theta$ 1-periodic is only completely monotone if $\theta$ is constant. I.e: a completely monotone function multiplied by a periodic function which isn't constant is not completely monotone.</p> <p>Now assume we have another solution $A(x)$ such that $a^{A(x)} = A(x+1)$, $A(0) = 1$ and $A'(x)$ is completely monotone. It follows with very little work, I'll leave it to the reader, that $A(x) = (^{x+\theta(x)}a)$ for some 1-periodic function $\theta$. So that $A'(x) = (1+\theta'(x))F(x+\theta(x))$. We also know that $1 + \theta'(x) \neq 0$ as this would destroy complete monotonicity. </p> <p>$A(x)$ can be written as</p> <p>$$A(x) = L + \sum_{n=1}^\infty a_n (-1)^n\lambda^{nx}\phi(x)^{n}$$</p> <p>where $\phi$ is one periodic. By just looking at this function it is rather clear its derivative can't be completely monotone. The $\phi$ is obviously going to kill the proper behaviour. To show this though, we'll need to use a good trick.</p> <p>For the moment, let $\mu = -\lambda^x \phi(x)$, then</p> <p>$$G(\mu) = L + \sum_{n=0}^\infty a_n \mu^n$$ </p> <p>This function is "fully monotonic," which is something that Sheldon showed in his post. The values $(-1)^na_n &lt; 0$ (kind of like the opposite of completely monotone). It is in fact the inverse Schroder function of $l(\mu) = a^\mu$. Where $l(G(\lambda \mu)) = G(\mu)$.</p> <p>Now, $-\mu = \lambda^x\phi(x)$ is not completely monotone (unless $\phi$ is constant), and therefore composing it with $G(\mu)$ breaks the monotonic structure. Therefore $\phi = C$ is constant. Therefore $A(x) = (^{x+C}a)$ but $A(0) = 1$, therefore $C = 0$ or is an imaginary period of $(^z a)$. Either way, $A(x) = (^xa)$.</p> <p>I think I can say, QED.</p> Answer 6: <p>I have written and submitted for publication <a href="http://tetration.org/IF.pdf" rel="noreferrer">The Existence and Uniqueness of the Taylor Series of Iterated Functions</a>. It has cleared an initial peer review with three knowledgeable mathematicians and now is in further peer review for publication. The intent is to have a single authoritative work on extending tetration and all the higher hyperoperators to the complex numbers. The uniqueness and existence theorems in the paper guarantee a unique solution to the problems. So at best, other extensions of tetration must be consistent with the extension provided in the paper. Basing the mathematics on iterated functions provides the generality needed to extend all the higher hyperoperators to complex numbers, not just tetration. </p> Answer 7: <p>Not sure if this is an answer but a good suggestion of how to show (1) and (2) (and perhaps an actual solution), and a good start for (3). Consider the Newton series formula for tetration. This series converges for all $\Re(z) &gt; -2$, and is very similar to your formula</p> <p>$$^za = \sum_{n=0}^\infty \sum_{k=0}^n (-1)^{n-k} \binom{z}{n}\binom{n}{k}(^k a)$$</p> <p>(The op claims this expression diverges, an alternate expression is</p> <p>$$\frac{1}{\Gamma(1-z)}\Big{(}\sum_{n=0}^\infty (^{n+1}a)\frac{(-1)^n}{n!(n+1-z)} + \int_1^\infty \big{(}\sum_{n=0}^\infty (^{n+1}a) \frac{(-x)^n}{n!}\big{)}x^{-z}\,dx\Big{)}$$</p> <p>of which the rest of this discussion still applies.)</p> <p>Now, since this solution is bounded in the half plane $\Re(z) &gt; 0$ it does have a uniqueness criterion--quite a good one at that too. It is the only solution to the functional equation to do such. In fact, if a solution $F$ exists where $F(z) = O(e^{\rho|\Re(z)| + \tau|\Im(z)|})$ for $\tau &lt; \pi/2$ then $F$ is still the above solution.</p> <p>What you have presented is a q-analog of the Newton series I just put above. Sadly I cannot find a rigorous paper detailing these types of series; I don't remember where I have seen them, but I have seen them. This suggests to me, that really you are just interpolating $(^na)$ in the same manner the Newton series does, except you are using q-analogs. It is necessary though that $q$ be fixed to the value you put (explanation below).</p> <p>Consider the rather beautiful fact that your solution is bounded in the right half plane. </p> <p>Observe if</p> <p>$[z]_q = \frac{1-q^z}{1-q}$, $[z]_q$ is periodic in $z$.</p> <p>$${z \brack n}_q = \frac{[z]_q!}{[n]_q![z-n]_q!} = \frac{[z]_q[z-1]_q...[z-n+1]_q}{[1]_q[2]_q...[n]_q}$$</p> <p>And each of these functions have a period, namely $2\pi i/\log(q)$..</p> <p>Therefore your final expansion has an imaginary period, it tends to $-W(\log(a))/\log(a)$ as $\Re(z) \to \infty$; thus, it is bounded in the right half plane. </p> <p>From this it follows your function is the standard tetration function (If it converges).</p> <p>So $t$ is bounded. The first obvious fact is $t(n) = (^n a)$. Next consider the wondrous power of a little known identity theorem for functions bounded in the right half plane. Namely, if $F(z)$ and $G(z)$ are bounded in the right half plane and $F\Big{|}_{\mathbb{N}} = G\Big{|}_{\mathbb{N}}$ then $F = G$. This follows because of Ramanujan's master theorem.</p> <p>I'll explain the identity theorem for a second, and then explain how this applies to your case. If $F$ and $G$ are bounded, then surely</p> <p>$$f(x) = \sum_{n=0}^\infty F(n+1)\frac{(-x)^n}{n!}$$ $$g(x) = \sum_{n=0}^\infty G(n+1)\frac{(-x)^n}{n!}$$</p> <p>are both entire functions, and by Ramanujan both satisfy</p> <p>$$\int_0^\infty f(x)x^{z-1}\,dx = \Gamma(z)F(1-z)$$ $$\int_0^\infty g(x)x^{z-1}\,dx = \Gamma(z)G(1-z)$$</p> <p>But $f = g$ and therefore $F = G$.</p> <p>Look at $t(z+1)$ and $a^{t(z)}$, both of these functions are bounded and agree on the naturals ($t(n+1) = a^{t(n)}$), therefore they are equal everywhere.</p> <p>Now as to the fact that this function is completely monotone: On the tetration forum Sheldon made a great point. I'm not sure of this fact, but I'm betting it's true. If $0 &lt; \lambda &lt; 1$ then $f(z) = \lambda^z$ for $\Re(z) &gt; 0$ is the only completely monotone function on $\mathbb{R}^+$ such that $\lambda f(z) = f(z+1)$. Another function would have to be $f(z+\theta(z))$ for $\theta$ a 1-periodic function, but this will most assuredly destroy the completely monotone structure. </p> <p>What's great about the standard solution (the one that is periodic, the one that is the Schroder iteration, the one that is bounded), is that it can be expanded in an exponential series for $\Re(z) &gt; 0$. This is just Fourier series, essentially.</p> <p>$$^za = F(z) = \sum_{k=0}^\infty a_k \lambda^{kz}$$</p> <p>$$F'(x) \sim C\lambda^x$$</p> <p>Since another solution is just $F(z+\theta(z))$ for $\theta$ periodic, we can most likely form a contradiction by using our knowledge of the uniqueness of the exponential function under these conditions. </p> <p>All in all, I mostly just had to weigh in my two cents, but I think all the conjectures you put are true and that your solution is in fact the 'standard' solution, which is: the periodic one, the bounded one, the Schroder one, and hopefully, the completely monotone one. Hopefully the rough proof I gave of (1) and (2) is good enough for you.</p> Answer 8: <p>Here is a proof sketch that the only completely monotone tetration is in fact the standard tetration. I am not certain, as I'm a little fuzzy on Sheldon's details on how to show tetrations derivative is completely monotone. As Sheldon wrote in the comments to my previous answer, the standard tetration (the Schroder iteration) is completely monotone. </p> <p>He shows it here <a href="https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129">https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129</a>. I think using similar tricks we can show this function is unique under these conditions. By all means my proof is not complete, but I think it's on the right track.</p> <p>First we start with the wondrous identity that</p> <p>$$^z a = L + \sum_{n=1}^\infty a_n(-1)^n \lambda^{nz}$$</p> <p>where $a_n(-1)^n &lt; 0$ for all $n$.</p> <p>Then if</p> <p>$$F(z) = \frac{d}{dz}(^za)$$</p> <p>it is completely monotone. We wish to show this is the sole function to satisfy this criterion.</p> <p>This will be our starting point. The next fact to consider is that $\lambda^{x + \theta(x)}$ for $\theta$ 1-periodic is only completely monotone if $\theta$ is constant. I.e: a completely monotone function multiplied by a periodic function which isn't constant is not completely monotone.</p> <p>Now assume we have another solution $A(x)$ such that $a^{A(x)} = A(x+1)$, $A(0) = 1$ and $A'(x)$ is completely monotone. It follows with very little work, I'll leave it to the reader, that $A(x) = (^{x+\theta(x)}a)$ for some 1-periodic function $\theta$. So that $A'(x) = (1+\theta'(x))F(x+\theta(x))$. We also know that $1 + \theta'(x) \neq 0$ as this would destroy complete monotonicity. </p> <p>$A(x)$ can be written as</p> <p>$$A(x) = L + \sum_{n=1}^\infty a_n (-1)^n\lambda^{nx}\phi(x)^{n}$$</p> <p>where $\phi$ is one periodic. By just looking at this function it is rather clear its derivative can't be completely monotone. The $\phi$ is obviously going to kill the proper behaviour. To show this though, we'll need to use a good trick.</p> <p>For the moment, let $\mu = -\lambda^x \phi(x)$, then</p> <p>$$G(\mu) = L + \sum_{n=0}^\infty a_n \mu^n$$ </p> <p>This function is "fully monotonic," which is something that Sheldon showed in his post. The values $(-1)^na_n &lt; 0$ (kind of like the opposite of completely monotone). It is in fact the inverse Schroder function of $l(\mu) = a^\mu$. Where $l(G(\lambda \mu)) = G(\mu)$.</p> <p>Now, $-\mu = \lambda^x\phi(x)$ is not completely monotone (unless $\phi$ is constant), and therefore composing it with $G(\mu)$ breaks the monotonic structure. Therefore $\phi = C$ is constant. Therefore $A(x) = (^{x+C}a)$ but $A(0) = 1$, therefore $C = 0$ or is an imaginary period of $(^z a)$. Either way, $A(x) = (^xa)$.</p> <p>I think I can say, QED.</p> Answer 9: <p>I have written and submitted for publication <a href="http://tetration.org/IF.pdf" rel="noreferrer">The Existence and Uniqueness of the Taylor Series of Iterated Functions</a>. It has cleared an initial peer review with three knowledgeable mathematicians and now is in further peer review for publication. The intent is to have a single authoritative work on extending tetration and all the higher hyperoperators to the complex numbers. The uniqueness and existence theorems in the paper guarantee a unique solution to the problems. So at best, other extensions of tetration must be consistent with the extension provided in the paper. Basing the mathematics on iterated functions provides the generality needed to extend all the higher hyperoperators to complex numbers, not just tetration. </p> Answer 10: <p>Not sure if this is an answer but a good suggestion of how to show (1) and (2) (and perhaps an actual solution), and a good start for (3). Consider the Newton series formula for tetration. This series converges for all $\Re(z) &gt; -2$, and is very similar to your formula</p> <p>$$^za = \sum_{n=0}^\infty \sum_{k=0}^n (-1)^{n-k} \binom{z}{n}\binom{n}{k}(^k a)$$</p> <p>(The op claims this expression diverges, an alternate expression is</p> <p>$$\frac{1}{\Gamma(1-z)}\Big{(}\sum_{n=0}^\infty (^{n+1}a)\frac{(-1)^n}{n!(n+1-z)} + \int_1^\infty \big{(}\sum_{n=0}^\infty (^{n+1}a) \frac{(-x)^n}{n!}\big{)}x^{-z}\,dx\Big{)}$$</p> <p>of which the rest of this discussion still applies.)</p> <p>Now, since this solution is bounded in the half plane $\Re(z) &gt; 0$ it does have a uniqueness criterion--quite a good one at that too. It is the only solution to the functional equation to do such. In fact, if a solution $F$ exists where $F(z) = O(e^{\rho|\Re(z)| + \tau|\Im(z)|})$ for $\tau &lt; \pi/2$ then $F$ is still the above solution.</p> <p>What you have presented is a q-analog of the Newton series I just put above. Sadly I cannot find a rigorous paper detailing these types of series; I don't remember where I have seen them, but I have seen them. This suggests to me, that really you are just interpolating $(^na)$ in the same manner the Newton series does, except you are using q-analogs. It is necessary though that $q$ be fixed to the value you put (explanation below).</p> <p>Consider the rather beautiful fact that your solution is bounded in the right half plane. </p> <p>Observe if</p> <p>$[z]_q = \frac{1-q^z}{1-q}$, $[z]_q$ is periodic in $z$.</p> <p>$${z \brack n}_q = \frac{[z]_q!}{[n]_q![z-n]_q!} = \frac{[z]_q[z-1]_q...[z-n+1]_q}{[1]_q[2]_q...[n]_q}$$</p> <p>And each of these functions have a period, namely $2\pi i/\log(q)$..</p> <p>Therefore your final expansion has an imaginary period, it tends to $-W(\log(a))/\log(a)$ as $\Re(z) \to \infty$; thus, it is bounded in the right half plane. </p> <p>From this it follows your function is the standard tetration function (If it converges).</p> <p>So $t$ is bounded. The first obvious fact is $t(n) = (^n a)$. Next consider the wondrous power of a little known identity theorem for functions bounded in the right half plane. Namely, if $F(z)$ and $G(z)$ are bounded in the right half plane and $F\Big{|}_{\mathbb{N}} = G\Big{|}_{\mathbb{N}}$ then $F = G$. This follows because of Ramanujan's master theorem.</p> <p>I'll explain the identity theorem for a second, and then explain how this applies to your case. If $F$ and $G$ are bounded, then surely</p> <p>$$f(x) = \sum_{n=0}^\infty F(n+1)\frac{(-x)^n}{n!}$$ $$g(x) = \sum_{n=0}^\infty G(n+1)\frac{(-x)^n}{n!}$$</p> <p>are both entire functions, and by Ramanujan both satisfy</p> <p>$$\int_0^\infty f(x)x^{z-1}\,dx = \Gamma(z)F(1-z)$$ $$\int_0^\infty g(x)x^{z-1}\,dx = \Gamma(z)G(1-z)$$</p> <p>But $f = g$ and therefore $F = G$.</p> <p>Look at $t(z+1)$ and $a^{t(z)}$, both of these functions are bounded and agree on the naturals ($t(n+1) = a^{t(n)}$), therefore they are equal everywhere.</p> <p>Now as to the fact that this function is completely monotone: On the tetration forum Sheldon made a great point. I'm not sure of this fact, but I'm betting it's true. If $0 &lt; \lambda &lt; 1$ then $f(z) = \lambda^z$ for $\Re(z) &gt; 0$ is the only completely monotone function on $\mathbb{R}^+$ such that $\lambda f(z) = f(z+1)$. Another function would have to be $f(z+\theta(z))$ for $\theta$ a 1-periodic function, but this will most assuredly destroy the completely monotone structure. </p> <p>What's great about the standard solution (the one that is periodic, the one that is the Schroder iteration, the one that is bounded), is that it can be expanded in an exponential series for $\Re(z) &gt; 0$. This is just Fourier series, essentially.</p> <p>$$^za = F(z) = \sum_{k=0}^\infty a_k \lambda^{kz}$$</p> <p>$$F'(x) \sim C\lambda^x$$</p> <p>Since another solution is just $F(z+\theta(z))$ for $\theta$ periodic, we can most likely form a contradiction by using our knowledge of the uniqueness of the exponential function under these conditions. </p> <p>All in all, I mostly just had to weigh in my two cents, but I think all the conjectures you put are true and that your solution is in fact the 'standard' solution, which is: the periodic one, the bounded one, the Schroder one, and hopefully, the completely monotone one. Hopefully the rough proof I gave of (1) and (2) is good enough for you.</p> Answer 11: <p>Here is a proof sketch that the only completely monotone tetration is in fact the standard tetration. I am not certain, as I'm a little fuzzy on Sheldon's details on how to show tetrations derivative is completely monotone. As Sheldon wrote in the comments to my previous answer, the standard tetration (the Schroder iteration) is completely monotone. </p> <p>He shows it here <a href="https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129">https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration/1996129#1996129</a>. I think using similar tricks we can show this function is unique under these conditions. By all means my proof is not complete, but I think it's on the right track.</p> <p>First we start with the wondrous identity that</p> <p>$$^z a = L + \sum_{n=1}^\infty a_n(-1)^n \lambda^{nz}$$</p> <p>where $a_n(-1)^n &lt; 0$ for all $n$.</p> <p>Then if</p> <p>$$F(z) = \frac{d}{dz}(^za)$$</p> <p>it is completely monotone. We wish to show this is the sole function to satisfy this criterion.</p> <p>This will be our starting point. The next fact to consider is that $\lambda^{x + \theta(x)}$ for $\theta$ 1-periodic is only completely monotone if $\theta$ is constant. I.e: a completely monotone function multiplied by a periodic function which isn't constant is not completely monotone.</p> <p>Now assume we have another solution $A(x)$ such that $a^{A(x)} = A(x+1)$, $A(0) = 1$ and $A'(x)$ is completely monotone. It follows with very little work, I'll leave it to the reader, that $A(x) = (^{x+\theta(x)}a)$ for some 1-periodic function $\theta$. So that $A'(x) = (1+\theta'(x))F(x+\theta(x))$. We also know that $1 + \theta'(x) \neq 0$ as this would destroy complete monotonicity. </p> <p>$A(x)$ can be written as</p> <p>$$A(x) = L + \sum_{n=1}^\infty a_n (-1)^n\lambda^{nx}\phi(x)^{n}$$</p> <p>where $\phi$ is one periodic. By just looking at this function it is rather clear its derivative can't be completely monotone. The $\phi$ is obviously going to kill the proper behaviour. To show this though, we'll need to use a good trick.</p> <p>For the moment, let $\mu = -\lambda^x \phi(x)$, then</p> <p>$$G(\mu) = L + \sum_{n=0}^\infty a_n \mu^n$$ </p> <p>This function is "fully monotonic," which is something that Sheldon showed in his post. The values $(-1)^na_n &lt; 0$ (kind of like the opposite of completely monotone). It is in fact the inverse Schroder function of $l(\mu) = a^\mu$. Where $l(G(\lambda \mu)) = G(\mu)$.</p> <p>Now, $-\mu = \lambda^x\phi(x)$ is not completely monotone (unless $\phi$ is constant), and therefore composing it with $G(\mu)$ breaks the monotonic structure. Therefore $\phi = C$ is constant. Therefore $A(x) = (^{x+C}a)$ but $A(0) = 1$, therefore $C = 0$ or is an imaginary period of $(^z a)$. Either way, $A(x) = (^xa)$.</p> <p>I think I can say, QED.</p> Answer 12: <p>I have written and submitted for publication <a href="http://tetration.org/IF.pdf" rel="noreferrer">The Existence and Uniqueness of the Taylor Series of Iterated Functions</a>. It has cleared an initial peer review with three knowledgeable mathematicians and now is in further peer review for publication. The intent is to have a single authoritative work on extending tetration and all the higher hyperoperators to the complex numbers. The uniqueness and existence theorems in the paper guarantee a unique solution to the problems. So at best, other extensions of tetration must be consistent with the extension provided in the paper. Basing the mathematics on iterated functions provides the generality needed to extend all the higher hyperoperators to complex numbers, not just tetration. </p>
Applications of microlocal analysis?
258,756
28
https://mathoverflow.net/questions/258756/applications-of-microlocal-analysis
<p>What examples are there of striking applications of the ideas of microlocal analysis?</p> <p>Ideally I'm looking for specific results in any of the relevant fields (PDE, algebraic/differential geometry/topology, representation theory etc.) which follow as natural consequences of microlocal analysis and are not known to be obtained easily without it.</p>
[ { "answer_id": 258758, "score": 24, "text": "<p>Microlocal analysis is used in <strong>computed tomography</strong> and other tomographic <strong>imaging techniques</strong> e.g. in <strong>medicine</strong> .</p>\n<p>Specifically, it is used to describe which wavefront sets (here: boundaries of objects, e.g. of organs in the human body) can or can not be detected by a specific tomographic measurement setup and also helps to understand reconstruction artifacts and develop strategies to overcome these. Details can be found for example in</p>\n<ul>\n<li>the overview article <a href=\"https://sites.tufts.edu/tquinto/files/2021/01//Wavefront-main.pdf\" rel=\"nofollow noreferrer\">Microlocal analysis in tomography</a> by Krishnan and Quinto and</li>\n<li>the &quot;Mathematics of Planet Earth&quot; article <a href=\"http://mpe.dimacs.rutgers.edu/2013/09/05/microlocal-analysis-and-imaging/\" rel=\"nofollow noreferrer\">Microlocal Analysis and Imaging</a>.</li>\n</ul>\n" }, { "answer_id": 259505, "score": 17, "text": "<p>Here are some examples:</p>\n\n<ol>\n<li>Helgason conjecture: Generically, the spherical principal series representations of a connected real semisimple Lie group $G$ are isomorphic via the Poisson transform to translation representations of $G$ on the spaces of joint eigenfunctions on the symmetric space $X=G/K$. The conjecture was proved in Kashiwara et al., Ann. Math. 107 (1978) using the boundary value theory in Kashiwara, Oshima, Ann. Math. 106 (1977). The inverse of the Poisson transform is shown to be a boundary value map defined by transforming, microlocally at the conormal bundle of a corner (Weyl chamber at infinity), the differential operators of the eigensystem on $X$ to simple normal form.</li>\n<li>Second term in Weyl asymptotics: The number $N(\\lambda)$ of eigenvalues $\\leq\\lambda$ for the Dirichlet problem in $\\Omega\\subset\\mathbb{R}^3$ of the Laplacian has the asymptotics $N(\\lambda)=c_1\\lambda^{3/2}+c_2\\lambda+o(\\lambda)$ if the closed geodesics cover only a set of measure zero in the unit sphere bundle. Up to constant factors $c_1$ and $c_2$ are the volumes of $\\Omega$ and of its boundary $\\partial\\Omega$. This was proved by Ivrii (1980); see Corollary 29.3.4 in volume 4 of Hörmander's treatise. The existence of a two-term asymptotics under certain assumptions had been a long-standing conjecture.</li>\n<li>Distribution of scattering poles: The Lax--Phillips scattering matrix for acoustic scattering by a non-trapping obstacle has no poles in a logarithmic neighbourhood of the real axis. This was conjectured by Lax and Phillips in their 1967 book (Conjecture (a) on page 155) and proved by Melrose and Sjöstrand, CPAM 35 (1982), as a corollary of their results on propagation of wavefront sets in boundary problems.</li>\n<li>Controllability in control theory: The solutions of a wave equation are controllable in time $T&gt;0$ from a subset $\\Gamma$ in the space-time boundary iff every geometric optics ray of time length $T$ intersects $\\Gamma$. This was shown in Bardos, Lebeau, Rauch, SIAM J. Control Opt. 30 (1992); the proof uses the results of Melrose and Sjöstrand mentioned above.</li>\n</ol>\n\n<p>The list can be continued with trace formulae, applications to inverse problems, and more. Many results have been developed further in semiclassical analysis, a modern variant of microlocal analysis, which works under weaker assumptions, and is applicable to Schrödinger's equation, for example.</p>\n" }, { "answer_id": 258768, "score": 16, "text": "<p>There are striking applications in dynamical systems, due to Dyatlov and Zworski, where dynamical zeta functions were analysed using techniques from microlocal analysis. These zeta functions have a long history in dynamical systems and microlocal analysis is giving a range of new tools. </p>\n" }, { "answer_id": 258775, "score": 15, "text": "<p>Although microlocal analysis was developed originally exclusively for linear problems, it has played an increasingly important role in nonlinear PDE via what's known as paradifferential calculus. There are some nice write-ups about this. Here are a few:</p>\n<p><a href=\"https://www.math.ucla.edu/%7Etao/247b.1.07w/notes6.pdf\" rel=\"nofollow noreferrer\">https://www.math.ucla.edu/~tao/247b.1.07w/notes6.pdf</a></p>\n<p><a href=\"http://www.ams.org/notices/201007/rtx100700858p.pdf\" rel=\"nofollow noreferrer\">http://www.ams.org/notices/201007/rtx100700858p.pdf</a></p>\n<p><a href=\"http://www.unc.edu/math/Faculty/met/micro2.pdf\" rel=\"nofollow noreferrer\">http://www.unc.edu/math/Faculty/met/micro2.pdf</a> (<a href=\"https://web.archive.org/web/20180114042916/http://www.unc.edu/math/Faculty/met/micro2.pdf\" rel=\"nofollow noreferrer\">Wayback Machine</a>)</p>\n" }, { "answer_id": 258893, "score": 12, "text": "<p>(1) The older and more widely known applications are to regularity and solvability of PDEs of any order that are <strong>not</strong> necessarily elliptic. </p>\n\n<p>(2) The phenomenon of the <em>propagation of singularities</em> along certain submanifolds of the cotangent space is also another striking application. </p>\n\n<p>These refined notions of <em>hypoellipticity</em> and <em>solvability</em> as well as the propagation of singularities involve many fantastic results which are possible because of microlocal analysis. Even the statements of many of these results <strong>require</strong> microlocal analysis. </p>\n\n<p>(3) There are also several applications to several complex variables and a very active subfield called CR (for Cauchy Riemann) manifolds. </p>\n\n<p>Some of the standard references on these topics are the two volumes of F. Treves on <em>Introduction to Pseudodifferential and Fourier Integral Operators</em> and the L. Hormander's treatise (Volumes I, II, III, etc).</p>\n" }, { "answer_id": 258810, "score": 5, "text": "<p>Search for papers by Maarten de Hoop (Colorado School of Mines, then Purdue, now a Simons chair at Rice University), who has used microlocal analysis very extensively to study problems in global and exploration seismology and imaging. </p>\n" }, { "answer_id": 260695, "score": 5, "text": "<p>Broadly speaking, microlocal analysis helps one in 'geometrization' of certain results on the singularities of distributions. In this direction, one striking application of microlocal analysis is the Hormander condition for the existence of product of distributions. Study of nonlinear pdes and the <a href=\"http://projecteuclid.org/download/pdf_1/euclid.cmp/1104287114\" rel=\"nofollow noreferrer\">renormalization problem in Quantum Field Theory</a> are two major applications of product of distributions.</p>\n<p>An important tool in this context is the wavefront set of a distribution. Roughly speaking, the wavefront set locates the singularities of a distribution in space and also helps in the analysis of the 'direction' of singularities. For <span class=\"math-container\">$u\\in \\mathcal{D}'(\\mathbb{R}^n)$</span> the wavefront set <span class=\"math-container\">$WF(u)\\subset T^{*}(\\mathbb{R}^n)$</span>. If <span class=\"math-container\">$(x,\\xi)\\in WF(u)$</span> then <span class=\"math-container\">$\\xi\\in \\mathbb{R}^{n}-\\{0\\}$</span> is the direction of 'propagation of singularity' that is located at <span class=\"math-container\">$x\\in\\mathbb{R}^{n}$</span>.</p>\n<p><strong>Hormander condition for existence of product:</strong> Let <span class=\"math-container\">$u,v\\in \\mathcal{D}'({\\mathbb{R}^n})$</span>. If for every <span class=\"math-container\">$(x,\\xi)\\in WF(u)$</span> we have <span class=\"math-container\">$(x,-\\xi)\\notin WF(v)$</span> then the product <span class=\"math-container\">$uv$</span> exists and the Leibniz rule <span class=\"math-container\">$\\partial(uv)=\\partial(u)v+u\\partial(v)$</span> holds. Moreover,\n<span class=\"math-container\">$WF(uv)\\subseteq WF(u)\\cup WF(v)\\cup \\{(x,\\xi+\\eta):(x,\\xi)\\in WF(u), (x,\\eta)\\in WF(v)\\}$</span>.</p>\n<p>We can note the following from the above condition:</p>\n<ol>\n<li>If the point of singularity is same for both distributions then for the product to exist the 'directions' of the singularity <em>should not cancel</em> each other.</li>\n<li>We can also estimate the wavefront set of the product.</li>\n<li>The condition is only a sufficiency condition. The condition only provides information on the product of distributions that are relevant for pdes - those that satisfy the Lebiniz rule. For example, the Heaviside function can be multiplied with itself but the Lebniz rule does not hold for the square of Heaviside function. So, the Hormander condition rules out square of Heaviside function.</li>\n</ol>\n" }, { "answer_id": 417384, "score": 5, "text": "<p>A pretty recent and impressive application of microlocal analysis is to analytic number theory or automorphic forms, via the <a href=\"https://arxiv.org/pdf/1805.07750.pdf\" rel=\"noreferrer\">orbit method as developed by Nelson and Venkatesh</a>. In <em>loc. cit.</em> they set the stage for the theory specifically thought to be applied to problems in the area of automorphic forms (trace formulas, moments, subconvexity, and more widely problems on large families of automorphic forms).</p>\n<p>A brilliant outcome of this setting is the proof of <a href=\"https://arxiv.org/pdf/2109.15230.pdf\" rel=\"noreferrer\">subconvexity for all L-functions</a> of general linear groups, by Nelson.</p>\n" }, { "answer_id": 258758, "score": 24, "text": "<p>Microlocal analysis is used in <strong>computed tomography</strong> and other tomographic <strong>imaging techniques</strong> e.g. in <strong>medicine</strong> .</p>\n<p>Specifically, it is used to describe which wavefront sets (here: boundaries of objects, e.g. of organs in the human body) can or can not be detected by a specific tomographic measurement setup and also helps to understand reconstruction artifacts and develop strategies to overcome these. Details can be found for example in</p>\n<ul>\n<li>the overview article <a href=\"https://sites.tufts.edu/tquinto/files/2021/01//Wavefront-main.pdf\" rel=\"nofollow noreferrer\">Microlocal analysis in tomography</a> by Krishnan and Quinto and</li>\n<li>the &quot;Mathematics of Planet Earth&quot; article <a href=\"http://mpe.dimacs.rutgers.edu/2013/09/05/microlocal-analysis-and-imaging/\" rel=\"nofollow noreferrer\">Microlocal Analysis and Imaging</a>.</li>\n</ul>\n" }, { "answer_id": 259505, "score": 17, "text": "<p>Here are some examples:</p>\n\n<ol>\n<li>Helgason conjecture: Generically, the spherical principal series representations of a connected real semisimple Lie group $G$ are isomorphic via the Poisson transform to translation representations of $G$ on the spaces of joint eigenfunctions on the symmetric space $X=G/K$. The conjecture was proved in Kashiwara et al., Ann. Math. 107 (1978) using the boundary value theory in Kashiwara, Oshima, Ann. Math. 106 (1977). The inverse of the Poisson transform is shown to be a boundary value map defined by transforming, microlocally at the conormal bundle of a corner (Weyl chamber at infinity), the differential operators of the eigensystem on $X$ to simple normal form.</li>\n<li>Second term in Weyl asymptotics: The number $N(\\lambda)$ of eigenvalues $\\leq\\lambda$ for the Dirichlet problem in $\\Omega\\subset\\mathbb{R}^3$ of the Laplacian has the asymptotics $N(\\lambda)=c_1\\lambda^{3/2}+c_2\\lambda+o(\\lambda)$ if the closed geodesics cover only a set of measure zero in the unit sphere bundle. Up to constant factors $c_1$ and $c_2$ are the volumes of $\\Omega$ and of its boundary $\\partial\\Omega$. This was proved by Ivrii (1980); see Corollary 29.3.4 in volume 4 of Hörmander's treatise. The existence of a two-term asymptotics under certain assumptions had been a long-standing conjecture.</li>\n<li>Distribution of scattering poles: The Lax--Phillips scattering matrix for acoustic scattering by a non-trapping obstacle has no poles in a logarithmic neighbourhood of the real axis. This was conjectured by Lax and Phillips in their 1967 book (Conjecture (a) on page 155) and proved by Melrose and Sjöstrand, CPAM 35 (1982), as a corollary of their results on propagation of wavefront sets in boundary problems.</li>\n<li>Controllability in control theory: The solutions of a wave equation are controllable in time $T&gt;0$ from a subset $\\Gamma$ in the space-time boundary iff every geometric optics ray of time length $T$ intersects $\\Gamma$. This was shown in Bardos, Lebeau, Rauch, SIAM J. Control Opt. 30 (1992); the proof uses the results of Melrose and Sjöstrand mentioned above.</li>\n</ol>\n\n<p>The list can be continued with trace formulae, applications to inverse problems, and more. Many results have been developed further in semiclassical analysis, a modern variant of microlocal analysis, which works under weaker assumptions, and is applicable to Schrödinger's equation, for example.</p>\n" }, { "answer_id": 258768, "score": 16, "text": "<p>There are striking applications in dynamical systems, due to Dyatlov and Zworski, where dynamical zeta functions were analysed using techniques from microlocal analysis. These zeta functions have a long history in dynamical systems and microlocal analysis is giving a range of new tools. </p>\n" }, { "answer_id": 258775, "score": 15, "text": "<p>Although microlocal analysis was developed originally exclusively for linear problems, it has played an increasingly important role in nonlinear PDE via what's known as paradifferential calculus. There are some nice write-ups about this. Here are a few:</p>\n<p><a href=\"https://www.math.ucla.edu/%7Etao/247b.1.07w/notes6.pdf\" rel=\"nofollow noreferrer\">https://www.math.ucla.edu/~tao/247b.1.07w/notes6.pdf</a></p>\n<p><a href=\"http://www.ams.org/notices/201007/rtx100700858p.pdf\" rel=\"nofollow noreferrer\">http://www.ams.org/notices/201007/rtx100700858p.pdf</a></p>\n<p><a href=\"http://www.unc.edu/math/Faculty/met/micro2.pdf\" rel=\"nofollow noreferrer\">http://www.unc.edu/math/Faculty/met/micro2.pdf</a> (<a href=\"https://web.archive.org/web/20180114042916/http://www.unc.edu/math/Faculty/met/micro2.pdf\" rel=\"nofollow noreferrer\">Wayback Machine</a>)</p>\n" }, { "answer_id": 258893, "score": 12, "text": "<p>(1) The older and more widely known applications are to regularity and solvability of PDEs of any order that are <strong>not</strong> necessarily elliptic. </p>\n\n<p>(2) The phenomenon of the <em>propagation of singularities</em> along certain submanifolds of the cotangent space is also another striking application. </p>\n\n<p>These refined notions of <em>hypoellipticity</em> and <em>solvability</em> as well as the propagation of singularities involve many fantastic results which are possible because of microlocal analysis. Even the statements of many of these results <strong>require</strong> microlocal analysis. </p>\n\n<p>(3) There are also several applications to several complex variables and a very active subfield called CR (for Cauchy Riemann) manifolds. </p>\n\n<p>Some of the standard references on these topics are the two volumes of F. Treves on <em>Introduction to Pseudodifferential and Fourier Integral Operators</em> and the L. Hormander's treatise (Volumes I, II, III, etc).</p>\n" }, { "answer_id": 258810, "score": 5, "text": "<p>Search for papers by Maarten de Hoop (Colorado School of Mines, then Purdue, now a Simons chair at Rice University), who has used microlocal analysis very extensively to study problems in global and exploration seismology and imaging. </p>\n" }, { "answer_id": 260695, "score": 5, "text": "<p>Broadly speaking, microlocal analysis helps one in 'geometrization' of certain results on the singularities of distributions. In this direction, one striking application of microlocal analysis is the Hormander condition for the existence of product of distributions. Study of nonlinear pdes and the <a href=\"http://projecteuclid.org/download/pdf_1/euclid.cmp/1104287114\" rel=\"nofollow noreferrer\">renormalization problem in Quantum Field Theory</a> are two major applications of product of distributions.</p>\n<p>An important tool in this context is the wavefront set of a distribution. Roughly speaking, the wavefront set locates the singularities of a distribution in space and also helps in the analysis of the 'direction' of singularities. For <span class=\"math-container\">$u\\in \\mathcal{D}'(\\mathbb{R}^n)$</span> the wavefront set <span class=\"math-container\">$WF(u)\\subset T^{*}(\\mathbb{R}^n)$</span>. If <span class=\"math-container\">$(x,\\xi)\\in WF(u)$</span> then <span class=\"math-container\">$\\xi\\in \\mathbb{R}^{n}-\\{0\\}$</span> is the direction of 'propagation of singularity' that is located at <span class=\"math-container\">$x\\in\\mathbb{R}^{n}$</span>.</p>\n<p><strong>Hormander condition for existence of product:</strong> Let <span class=\"math-container\">$u,v\\in \\mathcal{D}'({\\mathbb{R}^n})$</span>. If for every <span class=\"math-container\">$(x,\\xi)\\in WF(u)$</span> we have <span class=\"math-container\">$(x,-\\xi)\\notin WF(v)$</span> then the product <span class=\"math-container\">$uv$</span> exists and the Leibniz rule <span class=\"math-container\">$\\partial(uv)=\\partial(u)v+u\\partial(v)$</span> holds. Moreover,\n<span class=\"math-container\">$WF(uv)\\subseteq WF(u)\\cup WF(v)\\cup \\{(x,\\xi+\\eta):(x,\\xi)\\in WF(u), (x,\\eta)\\in WF(v)\\}$</span>.</p>\n<p>We can note the following from the above condition:</p>\n<ol>\n<li>If the point of singularity is same for both distributions then for the product to exist the 'directions' of the singularity <em>should not cancel</em> each other.</li>\n<li>We can also estimate the wavefront set of the product.</li>\n<li>The condition is only a sufficiency condition. The condition only provides information on the product of distributions that are relevant for pdes - those that satisfy the Lebiniz rule. For example, the Heaviside function can be multiplied with itself but the Lebniz rule does not hold for the square of Heaviside function. So, the Hormander condition rules out square of Heaviside function.</li>\n</ol>\n" }, { "answer_id": 417384, "score": 5, "text": "<p>A pretty recent and impressive application of microlocal analysis is to analytic number theory or automorphic forms, via the <a href=\"https://arxiv.org/pdf/1805.07750.pdf\" rel=\"noreferrer\">orbit method as developed by Nelson and Venkatesh</a>. In <em>loc. cit.</em> they set the stage for the theory specifically thought to be applied to problems in the area of automorphic forms (trace formulas, moments, subconvexity, and more widely problems on large families of automorphic forms).</p>\n<p>A brilliant outcome of this setting is the proof of <a href=\"https://arxiv.org/pdf/2109.15230.pdf\" rel=\"noreferrer\">subconvexity for all L-functions</a> of general linear groups, by Nelson.</p>\n" }, { "answer_id": 258758, "score": 24, "text": "<p>Microlocal analysis is used in <strong>computed tomography</strong> and other tomographic <strong>imaging techniques</strong> e.g. in <strong>medicine</strong> .</p>\n<p>Specifically, it is used to describe which wavefront sets (here: boundaries of objects, e.g. of organs in the human body) can or can not be detected by a specific tomographic measurement setup and also helps to understand reconstruction artifacts and develop strategies to overcome these. Details can be found for example in</p>\n<ul>\n<li>the overview article <a href=\"https://sites.tufts.edu/tquinto/files/2021/01//Wavefront-main.pdf\" rel=\"nofollow noreferrer\">Microlocal analysis in tomography</a> by Krishnan and Quinto and</li>\n<li>the &quot;Mathematics of Planet Earth&quot; article <a href=\"http://mpe.dimacs.rutgers.edu/2013/09/05/microlocal-analysis-and-imaging/\" rel=\"nofollow noreferrer\">Microlocal Analysis and Imaging</a>.</li>\n</ul>\n" }, { "answer_id": 259505, "score": 17, "text": "<p>Here are some examples:</p>\n\n<ol>\n<li>Helgason conjecture: Generically, the spherical principal series representations of a connected real semisimple Lie group $G$ are isomorphic via the Poisson transform to translation representations of $G$ on the spaces of joint eigenfunctions on the symmetric space $X=G/K$. The conjecture was proved in Kashiwara et al., Ann. Math. 107 (1978) using the boundary value theory in Kashiwara, Oshima, Ann. Math. 106 (1977). The inverse of the Poisson transform is shown to be a boundary value map defined by transforming, microlocally at the conormal bundle of a corner (Weyl chamber at infinity), the differential operators of the eigensystem on $X$ to simple normal form.</li>\n<li>Second term in Weyl asymptotics: The number $N(\\lambda)$ of eigenvalues $\\leq\\lambda$ for the Dirichlet problem in $\\Omega\\subset\\mathbb{R}^3$ of the Laplacian has the asymptotics $N(\\lambda)=c_1\\lambda^{3/2}+c_2\\lambda+o(\\lambda)$ if the closed geodesics cover only a set of measure zero in the unit sphere bundle. Up to constant factors $c_1$ and $c_2$ are the volumes of $\\Omega$ and of its boundary $\\partial\\Omega$. This was proved by Ivrii (1980); see Corollary 29.3.4 in volume 4 of Hörmander's treatise. The existence of a two-term asymptotics under certain assumptions had been a long-standing conjecture.</li>\n<li>Distribution of scattering poles: The Lax--Phillips scattering matrix for acoustic scattering by a non-trapping obstacle has no poles in a logarithmic neighbourhood of the real axis. This was conjectured by Lax and Phillips in their 1967 book (Conjecture (a) on page 155) and proved by Melrose and Sjöstrand, CPAM 35 (1982), as a corollary of their results on propagation of wavefront sets in boundary problems.</li>\n<li>Controllability in control theory: The solutions of a wave equation are controllable in time $T&gt;0$ from a subset $\\Gamma$ in the space-time boundary iff every geometric optics ray of time length $T$ intersects $\\Gamma$. This was shown in Bardos, Lebeau, Rauch, SIAM J. Control Opt. 30 (1992); the proof uses the results of Melrose and Sjöstrand mentioned above.</li>\n</ol>\n\n<p>The list can be continued with trace formulae, applications to inverse problems, and more. Many results have been developed further in semiclassical analysis, a modern variant of microlocal analysis, which works under weaker assumptions, and is applicable to Schrödinger's equation, for example.</p>\n" }, { "answer_id": 258768, "score": 16, "text": "<p>There are striking applications in dynamical systems, due to Dyatlov and Zworski, where dynamical zeta functions were analysed using techniques from microlocal analysis. These zeta functions have a long history in dynamical systems and microlocal analysis is giving a range of new tools. </p>\n" }, { "answer_id": 258775, "score": 15, "text": "<p>Although microlocal analysis was developed originally exclusively for linear problems, it has played an increasingly important role in nonlinear PDE via what's known as paradifferential calculus. There are some nice write-ups about this. Here are a few:</p>\n<p><a href=\"https://www.math.ucla.edu/%7Etao/247b.1.07w/notes6.pdf\" rel=\"nofollow noreferrer\">https://www.math.ucla.edu/~tao/247b.1.07w/notes6.pdf</a></p>\n<p><a href=\"http://www.ams.org/notices/201007/rtx100700858p.pdf\" rel=\"nofollow noreferrer\">http://www.ams.org/notices/201007/rtx100700858p.pdf</a></p>\n<p><a href=\"http://www.unc.edu/math/Faculty/met/micro2.pdf\" rel=\"nofollow noreferrer\">http://www.unc.edu/math/Faculty/met/micro2.pdf</a> (<a href=\"https://web.archive.org/web/20180114042916/http://www.unc.edu/math/Faculty/met/micro2.pdf\" rel=\"nofollow noreferrer\">Wayback Machine</a>)</p>\n" }, { "answer_id": 258893, "score": 12, "text": "<p>(1) The older and more widely known applications are to regularity and solvability of PDEs of any order that are <strong>not</strong> necessarily elliptic. </p>\n\n<p>(2) The phenomenon of the <em>propagation of singularities</em> along certain submanifolds of the cotangent space is also another striking application. </p>\n\n<p>These refined notions of <em>hypoellipticity</em> and <em>solvability</em> as well as the propagation of singularities involve many fantastic results which are possible because of microlocal analysis. Even the statements of many of these results <strong>require</strong> microlocal analysis. </p>\n\n<p>(3) There are also several applications to several complex variables and a very active subfield called CR (for Cauchy Riemann) manifolds. </p>\n\n<p>Some of the standard references on these topics are the two volumes of F. Treves on <em>Introduction to Pseudodifferential and Fourier Integral Operators</em> and the L. Hormander's treatise (Volumes I, II, III, etc).</p>\n" }, { "answer_id": 258810, "score": 5, "text": "<p>Search for papers by Maarten de Hoop (Colorado School of Mines, then Purdue, now a Simons chair at Rice University), who has used microlocal analysis very extensively to study problems in global and exploration seismology and imaging. </p>\n" }, { "answer_id": 260695, "score": 5, "text": "<p>Broadly speaking, microlocal analysis helps one in 'geometrization' of certain results on the singularities of distributions. In this direction, one striking application of microlocal analysis is the Hormander condition for the existence of product of distributions. Study of nonlinear pdes and the <a href=\"http://projecteuclid.org/download/pdf_1/euclid.cmp/1104287114\" rel=\"nofollow noreferrer\">renormalization problem in Quantum Field Theory</a> are two major applications of product of distributions.</p>\n<p>An important tool in this context is the wavefront set of a distribution. Roughly speaking, the wavefront set locates the singularities of a distribution in space and also helps in the analysis of the 'direction' of singularities. For <span class=\"math-container\">$u\\in \\mathcal{D}'(\\mathbb{R}^n)$</span> the wavefront set <span class=\"math-container\">$WF(u)\\subset T^{*}(\\mathbb{R}^n)$</span>. If <span class=\"math-container\">$(x,\\xi)\\in WF(u)$</span> then <span class=\"math-container\">$\\xi\\in \\mathbb{R}^{n}-\\{0\\}$</span> is the direction of 'propagation of singularity' that is located at <span class=\"math-container\">$x\\in\\mathbb{R}^{n}$</span>.</p>\n<p><strong>Hormander condition for existence of product:</strong> Let <span class=\"math-container\">$u,v\\in \\mathcal{D}'({\\mathbb{R}^n})$</span>. If for every <span class=\"math-container\">$(x,\\xi)\\in WF(u)$</span> we have <span class=\"math-container\">$(x,-\\xi)\\notin WF(v)$</span> then the product <span class=\"math-container\">$uv$</span> exists and the Leibniz rule <span class=\"math-container\">$\\partial(uv)=\\partial(u)v+u\\partial(v)$</span> holds. Moreover,\n<span class=\"math-container\">$WF(uv)\\subseteq WF(u)\\cup WF(v)\\cup \\{(x,\\xi+\\eta):(x,\\xi)\\in WF(u), (x,\\eta)\\in WF(v)\\}$</span>.</p>\n<p>We can note the following from the above condition:</p>\n<ol>\n<li>If the point of singularity is same for both distributions then for the product to exist the 'directions' of the singularity <em>should not cancel</em> each other.</li>\n<li>We can also estimate the wavefront set of the product.</li>\n<li>The condition is only a sufficiency condition. The condition only provides information on the product of distributions that are relevant for pdes - those that satisfy the Lebiniz rule. For example, the Heaviside function can be multiplied with itself but the Lebniz rule does not hold for the square of Heaviside function. So, the Hormander condition rules out square of Heaviside function.</li>\n</ol>\n" }, { "answer_id": 417384, "score": 5, "text": "<p>A pretty recent and impressive application of microlocal analysis is to analytic number theory or automorphic forms, via the <a href=\"https://arxiv.org/pdf/1805.07750.pdf\" rel=\"noreferrer\">orbit method as developed by Nelson and Venkatesh</a>. In <em>loc. cit.</em> they set the stage for the theory specifically thought to be applied to problems in the area of automorphic forms (trace formulas, moments, subconvexity, and more widely problems on large families of automorphic forms).</p>\n<p>A brilliant outcome of this setting is the proof of <a href=\"https://arxiv.org/pdf/2109.15230.pdf\" rel=\"noreferrer\">subconvexity for all L-functions</a> of general linear groups, by Nelson.</p>\n" }, { "answer_id": 258758, "score": 24, "text": "<p>Microlocal analysis is used in <strong>computed tomography</strong> and other tomographic <strong>imaging techniques</strong> e.g. in <strong>medicine</strong> .</p>\n<p>Specifically, it is used to describe which wavefront sets (here: boundaries of objects, e.g. of organs in the human body) can or can not be detected by a specific tomographic measurement setup and also helps to understand reconstruction artifacts and develop strategies to overcome these. Details can be found for example in</p>\n<ul>\n<li>the overview article <a href=\"https://sites.tufts.edu/tquinto/files/2021/01//Wavefront-main.pdf\" rel=\"nofollow noreferrer\">Microlocal analysis in tomography</a> by Krishnan and Quinto and</li>\n<li>the &quot;Mathematics of Planet Earth&quot; article <a href=\"http://mpe.dimacs.rutgers.edu/2013/09/05/microlocal-analysis-and-imaging/\" rel=\"nofollow noreferrer\">Microlocal Analysis and Imaging</a>.</li>\n</ul>\n" }, { "answer_id": 259505, "score": 17, "text": "<p>Here are some examples:</p>\n\n<ol>\n<li>Helgason conjecture: Generically, the spherical principal series representations of a connected real semisimple Lie group $G$ are isomorphic via the Poisson transform to translation representations of $G$ on the spaces of joint eigenfunctions on the symmetric space $X=G/K$. The conjecture was proved in Kashiwara et al., Ann. Math. 107 (1978) using the boundary value theory in Kashiwara, Oshima, Ann. Math. 106 (1977). The inverse of the Poisson transform is shown to be a boundary value map defined by transforming, microlocally at the conormal bundle of a corner (Weyl chamber at infinity), the differential operators of the eigensystem on $X$ to simple normal form.</li>\n<li>Second term in Weyl asymptotics: The number $N(\\lambda)$ of eigenvalues $\\leq\\lambda$ for the Dirichlet problem in $\\Omega\\subset\\mathbb{R}^3$ of the Laplacian has the asymptotics $N(\\lambda)=c_1\\lambda^{3/2}+c_2\\lambda+o(\\lambda)$ if the closed geodesics cover only a set of measure zero in the unit sphere bundle. Up to constant factors $c_1$ and $c_2$ are the volumes of $\\Omega$ and of its boundary $\\partial\\Omega$. This was proved by Ivrii (1980); see Corollary 29.3.4 in volume 4 of Hörmander's treatise. The existence of a two-term asymptotics under certain assumptions had been a long-standing conjecture.</li>\n<li>Distribution of scattering poles: The Lax--Phillips scattering matrix for acoustic scattering by a non-trapping obstacle has no poles in a logarithmic neighbourhood of the real axis. This was conjectured by Lax and Phillips in their 1967 book (Conjecture (a) on page 155) and proved by Melrose and Sjöstrand, CPAM 35 (1982), as a corollary of their results on propagation of wavefront sets in boundary problems.</li>\n<li>Controllability in control theory: The solutions of a wave equation are controllable in time $T&gt;0$ from a subset $\\Gamma$ in the space-time boundary iff every geometric optics ray of time length $T$ intersects $\\Gamma$. This was shown in Bardos, Lebeau, Rauch, SIAM J. Control Opt. 30 (1992); the proof uses the results of Melrose and Sjöstrand mentioned above.</li>\n</ol>\n\n<p>The list can be continued with trace formulae, applications to inverse problems, and more. Many results have been developed further in semiclassical analysis, a modern variant of microlocal analysis, which works under weaker assumptions, and is applicable to Schrödinger's equation, for example.</p>\n" }, { "answer_id": 258768, "score": 16, "text": "<p>There are striking applications in dynamical systems, due to Dyatlov and Zworski, where dynamical zeta functions were analysed using techniques from microlocal analysis. These zeta functions have a long history in dynamical systems and microlocal analysis is giving a range of new tools. </p>\n" }, { "answer_id": 258775, "score": 15, "text": "<p>Although microlocal analysis was developed originally exclusively for linear problems, it has played an increasingly important role in nonlinear PDE via what's known as paradifferential calculus. There are some nice write-ups about this. Here are a few:</p>\n<p><a href=\"https://www.math.ucla.edu/%7Etao/247b.1.07w/notes6.pdf\" rel=\"nofollow noreferrer\">https://www.math.ucla.edu/~tao/247b.1.07w/notes6.pdf</a></p>\n<p><a href=\"http://www.ams.org/notices/201007/rtx100700858p.pdf\" rel=\"nofollow noreferrer\">http://www.ams.org/notices/201007/rtx100700858p.pdf</a></p>\n<p><a href=\"http://www.unc.edu/math/Faculty/met/micro2.pdf\" rel=\"nofollow noreferrer\">http://www.unc.edu/math/Faculty/met/micro2.pdf</a> (<a href=\"https://web.archive.org/web/20180114042916/http://www.unc.edu/math/Faculty/met/micro2.pdf\" rel=\"nofollow noreferrer\">Wayback Machine</a>)</p>\n" }, { "answer_id": 258893, "score": 12, "text": "<p>(1) The older and more widely known applications are to regularity and solvability of PDEs of any order that are <strong>not</strong> necessarily elliptic. </p>\n\n<p>(2) The phenomenon of the <em>propagation of singularities</em> along certain submanifolds of the cotangent space is also another striking application. </p>\n\n<p>These refined notions of <em>hypoellipticity</em> and <em>solvability</em> as well as the propagation of singularities involve many fantastic results which are possible because of microlocal analysis. Even the statements of many of these results <strong>require</strong> microlocal analysis. </p>\n\n<p>(3) There are also several applications to several complex variables and a very active subfield called CR (for Cauchy Riemann) manifolds. </p>\n\n<p>Some of the standard references on these topics are the two volumes of F. Treves on <em>Introduction to Pseudodifferential and Fourier Integral Operators</em> and the L. Hormander's treatise (Volumes I, II, III, etc).</p>\n" }, { "answer_id": 258810, "score": 5, "text": "<p>Search for papers by Maarten de Hoop (Colorado School of Mines, then Purdue, now a Simons chair at Rice University), who has used microlocal analysis very extensively to study problems in global and exploration seismology and imaging. </p>\n" }, { "answer_id": 260695, "score": 5, "text": "<p>Broadly speaking, microlocal analysis helps one in 'geometrization' of certain results on the singularities of distributions. In this direction, one striking application of microlocal analysis is the Hormander condition for the existence of product of distributions. Study of nonlinear pdes and the <a href=\"http://projecteuclid.org/download/pdf_1/euclid.cmp/1104287114\" rel=\"nofollow noreferrer\">renormalization problem in Quantum Field Theory</a> are two major applications of product of distributions.</p>\n<p>An important tool in this context is the wavefront set of a distribution. Roughly speaking, the wavefront set locates the singularities of a distribution in space and also helps in the analysis of the 'direction' of singularities. For <span class=\"math-container\">$u\\in \\mathcal{D}'(\\mathbb{R}^n)$</span> the wavefront set <span class=\"math-container\">$WF(u)\\subset T^{*}(\\mathbb{R}^n)$</span>. If <span class=\"math-container\">$(x,\\xi)\\in WF(u)$</span> then <span class=\"math-container\">$\\xi\\in \\mathbb{R}^{n}-\\{0\\}$</span> is the direction of 'propagation of singularity' that is located at <span class=\"math-container\">$x\\in\\mathbb{R}^{n}$</span>.</p>\n<p><strong>Hormander condition for existence of product:</strong> Let <span class=\"math-container\">$u,v\\in \\mathcal{D}'({\\mathbb{R}^n})$</span>. If for every <span class=\"math-container\">$(x,\\xi)\\in WF(u)$</span> we have <span class=\"math-container\">$(x,-\\xi)\\notin WF(v)$</span> then the product <span class=\"math-container\">$uv$</span> exists and the Leibniz rule <span class=\"math-container\">$\\partial(uv)=\\partial(u)v+u\\partial(v)$</span> holds. Moreover,\n<span class=\"math-container\">$WF(uv)\\subseteq WF(u)\\cup WF(v)\\cup \\{(x,\\xi+\\eta):(x,\\xi)\\in WF(u), (x,\\eta)\\in WF(v)\\}$</span>.</p>\n<p>We can note the following from the above condition:</p>\n<ol>\n<li>If the point of singularity is same for both distributions then for the product to exist the 'directions' of the singularity <em>should not cancel</em> each other.</li>\n<li>We can also estimate the wavefront set of the product.</li>\n<li>The condition is only a sufficiency condition. The condition only provides information on the product of distributions that are relevant for pdes - those that satisfy the Lebiniz rule. For example, the Heaviside function can be multiplied with itself but the Lebniz rule does not hold for the square of Heaviside function. So, the Hormander condition rules out square of Heaviside function.</li>\n</ol>\n" }, { "answer_id": 417384, "score": 5, "text": "<p>A pretty recent and impressive application of microlocal analysis is to analytic number theory or automorphic forms, via the <a href=\"https://arxiv.org/pdf/1805.07750.pdf\" rel=\"noreferrer\">orbit method as developed by Nelson and Venkatesh</a>. In <em>loc. cit.</em> they set the stage for the theory specifically thought to be applied to problems in the area of automorphic forms (trace formulas, moments, subconvexity, and more widely problems on large families of automorphic forms).</p>\n<p>A brilliant outcome of this setting is the proof of <a href=\"https://arxiv.org/pdf/2109.15230.pdf\" rel=\"noreferrer\">subconvexity for all L-functions</a> of general linear groups, by Nelson.</p>\n" } ]
Question: Applications of microlocal analysis? Answer 1: <p>Microlocal analysis is used in <strong>computed tomography</strong> and other tomographic <strong>imaging techniques</strong> e.g. in <strong>medicine</strong> .</p> <p>Specifically, it is used to describe which wavefront sets (here: boundaries of objects, e.g. of organs in the human body) can or can not be detected by a specific tomographic measurement setup and also helps to understand reconstruction artifacts and develop strategies to overcome these. Details can be found for example in</p> <ul> <li>the overview article <a href="https://sites.tufts.edu/tquinto/files/2021/01//Wavefront-main.pdf" rel="nofollow noreferrer">Microlocal analysis in tomography</a> by Krishnan and Quinto and</li> <li>the &quot;Mathematics of Planet Earth&quot; article <a href="http://mpe.dimacs.rutgers.edu/2013/09/05/microlocal-analysis-and-imaging/" rel="nofollow noreferrer">Microlocal Analysis and Imaging</a>.</li> </ul> Answer 2: <p>Here are some examples:</p> <ol> <li>Helgason conjecture: Generically, the spherical principal series representations of a connected real semisimple Lie group $G$ are isomorphic via the Poisson transform to translation representations of $G$ on the spaces of joint eigenfunctions on the symmetric space $X=G/K$. The conjecture was proved in Kashiwara et al., Ann. Math. 107 (1978) using the boundary value theory in Kashiwara, Oshima, Ann. Math. 106 (1977). The inverse of the Poisson transform is shown to be a boundary value map defined by transforming, microlocally at the conormal bundle of a corner (Weyl chamber at infinity), the differential operators of the eigensystem on $X$ to simple normal form.</li> <li>Second term in Weyl asymptotics: The number $N(\lambda)$ of eigenvalues $\leq\lambda$ for the Dirichlet problem in $\Omega\subset\mathbb{R}^3$ of the Laplacian has the asymptotics $N(\lambda)=c_1\lambda^{3/2}+c_2\lambda+o(\lambda)$ if the closed geodesics cover only a set of measure zero in the unit sphere bundle. Up to constant factors $c_1$ and $c_2$ are the volumes of $\Omega$ and of its boundary $\partial\Omega$. This was proved by Ivrii (1980); see Corollary 29.3.4 in volume 4 of Hörmander's treatise. The existence of a two-term asymptotics under certain assumptions had been a long-standing conjecture.</li> <li>Distribution of scattering poles: The Lax--Phillips scattering matrix for acoustic scattering by a non-trapping obstacle has no poles in a logarithmic neighbourhood of the real axis. This was conjectured by Lax and Phillips in their 1967 book (Conjecture (a) on page 155) and proved by Melrose and Sjöstrand, CPAM 35 (1982), as a corollary of their results on propagation of wavefront sets in boundary problems.</li> <li>Controllability in control theory: The solutions of a wave equation are controllable in time $T&gt;0$ from a subset $\Gamma$ in the space-time boundary iff every geometric optics ray of time length $T$ intersects $\Gamma$. This was shown in Bardos, Lebeau, Rauch, SIAM J. Control Opt. 30 (1992); the proof uses the results of Melrose and Sjöstrand mentioned above.</li> </ol> <p>The list can be continued with trace formulae, applications to inverse problems, and more. Many results have been developed further in semiclassical analysis, a modern variant of microlocal analysis, which works under weaker assumptions, and is applicable to Schrödinger's equation, for example.</p> Answer 3: <p>There are striking applications in dynamical systems, due to Dyatlov and Zworski, where dynamical zeta functions were analysed using techniques from microlocal analysis. These zeta functions have a long history in dynamical systems and microlocal analysis is giving a range of new tools. </p> Answer 4: <p>Although microlocal analysis was developed originally exclusively for linear problems, it has played an increasingly important role in nonlinear PDE via what's known as paradifferential calculus. There are some nice write-ups about this. Here are a few:</p> <p><a href="https://www.math.ucla.edu/%7Etao/247b.1.07w/notes6.pdf" rel="nofollow noreferrer">https://www.math.ucla.edu/~tao/247b.1.07w/notes6.pdf</a></p> <p><a href="http://www.ams.org/notices/201007/rtx100700858p.pdf" rel="nofollow noreferrer">http://www.ams.org/notices/201007/rtx100700858p.pdf</a></p> <p><a href="http://www.unc.edu/math/Faculty/met/micro2.pdf" rel="nofollow noreferrer">http://www.unc.edu/math/Faculty/met/micro2.pdf</a> (<a href="https://web.archive.org/web/20180114042916/http://www.unc.edu/math/Faculty/met/micro2.pdf" rel="nofollow noreferrer">Wayback Machine</a>)</p> Answer 5: <p>(1) The older and more widely known applications are to regularity and solvability of PDEs of any order that are <strong>not</strong> necessarily elliptic. </p> <p>(2) The phenomenon of the <em>propagation of singularities</em> along certain submanifolds of the cotangent space is also another striking application. </p> <p>These refined notions of <em>hypoellipticity</em> and <em>solvability</em> as well as the propagation of singularities involve many fantastic results which are possible because of microlocal analysis. Even the statements of many of these results <strong>require</strong> microlocal analysis. </p> <p>(3) There are also several applications to several complex variables and a very active subfield called CR (for Cauchy Riemann) manifolds. </p> <p>Some of the standard references on these topics are the two volumes of F. Treves on <em>Introduction to Pseudodifferential and Fourier Integral Operators</em> and the L. Hormander's treatise (Volumes I, II, III, etc).</p> Answer 6: <p>Search for papers by Maarten de Hoop (Colorado School of Mines, then Purdue, now a Simons chair at Rice University), who has used microlocal analysis very extensively to study problems in global and exploration seismology and imaging. </p> Answer 7: <p>Broadly speaking, microlocal analysis helps one in 'geometrization' of certain results on the singularities of distributions. In this direction, one striking application of microlocal analysis is the Hormander condition for the existence of product of distributions. Study of nonlinear pdes and the <a href="http://projecteuclid.org/download/pdf_1/euclid.cmp/1104287114" rel="nofollow noreferrer">renormalization problem in Quantum Field Theory</a> are two major applications of product of distributions.</p> <p>An important tool in this context is the wavefront set of a distribution. Roughly speaking, the wavefront set locates the singularities of a distribution in space and also helps in the analysis of the 'direction' of singularities. For <span class="math-container">$u\in \mathcal{D}'(\mathbb{R}^n)$</span> the wavefront set <span class="math-container">$WF(u)\subset T^{*}(\mathbb{R}^n)$</span>. If <span class="math-container">$(x,\xi)\in WF(u)$</span> then <span class="math-container">$\xi\in \mathbb{R}^{n}-\{0\}$</span> is the direction of 'propagation of singularity' that is located at <span class="math-container">$x\in\mathbb{R}^{n}$</span>.</p> <p><strong>Hormander condition for existence of product:</strong> Let <span class="math-container">$u,v\in \mathcal{D}'({\mathbb{R}^n})$</span>. If for every <span class="math-container">$(x,\xi)\in WF(u)$</span> we have <span class="math-container">$(x,-\xi)\notin WF(v)$</span> then the product <span class="math-container">$uv$</span> exists and the Leibniz rule <span class="math-container">$\partial(uv)=\partial(u)v+u\partial(v)$</span> holds. Moreover, <span class="math-container">$WF(uv)\subseteq WF(u)\cup WF(v)\cup \{(x,\xi+\eta):(x,\xi)\in WF(u), (x,\eta)\in WF(v)\}$</span>.</p> <p>We can note the following from the above condition:</p> <ol> <li>If the point of singularity is same for both distributions then for the product to exist the 'directions' of the singularity <em>should not cancel</em> each other.</li> <li>We can also estimate the wavefront set of the product.</li> <li>The condition is only a sufficiency condition. The condition only provides information on the product of distributions that are relevant for pdes - those that satisfy the Lebiniz rule. For example, the Heaviside function can be multiplied with itself but the Lebniz rule does not hold for the square of Heaviside function. So, the Hormander condition rules out square of Heaviside function.</li> </ol> Answer 8: <p>A pretty recent and impressive application of microlocal analysis is to analytic number theory or automorphic forms, via the <a href="https://arxiv.org/pdf/1805.07750.pdf" rel="noreferrer">orbit method as developed by Nelson and Venkatesh</a>. In <em>loc. cit.</em> they set the stage for the theory specifically thought to be applied to problems in the area of automorphic forms (trace formulas, moments, subconvexity, and more widely problems on large families of automorphic forms).</p> <p>A brilliant outcome of this setting is the proof of <a href="https://arxiv.org/pdf/2109.15230.pdf" rel="noreferrer">subconvexity for all L-functions</a> of general linear groups, by Nelson.</p> Answer 9: <p>Microlocal analysis is used in <strong>computed tomography</strong> and other tomographic <strong>imaging techniques</strong> e.g. in <strong>medicine</strong> .</p> <p>Specifically, it is used to describe which wavefront sets (here: boundaries of objects, e.g. of organs in the human body) can or can not be detected by a specific tomographic measurement setup and also helps to understand reconstruction artifacts and develop strategies to overcome these. Details can be found for example in</p> <ul> <li>the overview article <a href="https://sites.tufts.edu/tquinto/files/2021/01//Wavefront-main.pdf" rel="nofollow noreferrer">Microlocal analysis in tomography</a> by Krishnan and Quinto and</li> <li>the &quot;Mathematics of Planet Earth&quot; article <a href="http://mpe.dimacs.rutgers.edu/2013/09/05/microlocal-analysis-and-imaging/" rel="nofollow noreferrer">Microlocal Analysis and Imaging</a>.</li> </ul> Answer 10: <p>Here are some examples:</p> <ol> <li>Helgason conjecture: Generically, the spherical principal series representations of a connected real semisimple Lie group $G$ are isomorphic via the Poisson transform to translation representations of $G$ on the spaces of joint eigenfunctions on the symmetric space $X=G/K$. The conjecture was proved in Kashiwara et al., Ann. Math. 107 (1978) using the boundary value theory in Kashiwara, Oshima, Ann. Math. 106 (1977). The inverse of the Poisson transform is shown to be a boundary value map defined by transforming, microlocally at the conormal bundle of a corner (Weyl chamber at infinity), the differential operators of the eigensystem on $X$ to simple normal form.</li> <li>Second term in Weyl asymptotics: The number $N(\lambda)$ of eigenvalues $\leq\lambda$ for the Dirichlet problem in $\Omega\subset\mathbb{R}^3$ of the Laplacian has the asymptotics $N(\lambda)=c_1\lambda^{3/2}+c_2\lambda+o(\lambda)$ if the closed geodesics cover only a set of measure zero in the unit sphere bundle. Up to constant factors $c_1$ and $c_2$ are the volumes of $\Omega$ and of its boundary $\partial\Omega$. This was proved by Ivrii (1980); see Corollary 29.3.4 in volume 4 of Hörmander's treatise. The existence of a two-term asymptotics under certain assumptions had been a long-standing conjecture.</li> <li>Distribution of scattering poles: The Lax--Phillips scattering matrix for acoustic scattering by a non-trapping obstacle has no poles in a logarithmic neighbourhood of the real axis. This was conjectured by Lax and Phillips in their 1967 book (Conjecture (a) on page 155) and proved by Melrose and Sjöstrand, CPAM 35 (1982), as a corollary of their results on propagation of wavefront sets in boundary problems.</li> <li>Controllability in control theory: The solutions of a wave equation are controllable in time $T&gt;0$ from a subset $\Gamma$ in the space-time boundary iff every geometric optics ray of time length $T$ intersects $\Gamma$. This was shown in Bardos, Lebeau, Rauch, SIAM J. Control Opt. 30 (1992); the proof uses the results of Melrose and Sjöstrand mentioned above.</li> </ol> <p>The list can be continued with trace formulae, applications to inverse problems, and more. Many results have been developed further in semiclassical analysis, a modern variant of microlocal analysis, which works under weaker assumptions, and is applicable to Schrödinger's equation, for example.</p> Answer 11: <p>There are striking applications in dynamical systems, due to Dyatlov and Zworski, where dynamical zeta functions were analysed using techniques from microlocal analysis. These zeta functions have a long history in dynamical systems and microlocal analysis is giving a range of new tools. </p> Answer 12: <p>Although microlocal analysis was developed originally exclusively for linear problems, it has played an increasingly important role in nonlinear PDE via what's known as paradifferential calculus. There are some nice write-ups about this. Here are a few:</p> <p><a href="https://www.math.ucla.edu/%7Etao/247b.1.07w/notes6.pdf" rel="nofollow noreferrer">https://www.math.ucla.edu/~tao/247b.1.07w/notes6.pdf</a></p> <p><a href="http://www.ams.org/notices/201007/rtx100700858p.pdf" rel="nofollow noreferrer">http://www.ams.org/notices/201007/rtx100700858p.pdf</a></p> <p><a href="http://www.unc.edu/math/Faculty/met/micro2.pdf" rel="nofollow noreferrer">http://www.unc.edu/math/Faculty/met/micro2.pdf</a> (<a href="https://web.archive.org/web/20180114042916/http://www.unc.edu/math/Faculty/met/micro2.pdf" rel="nofollow noreferrer">Wayback Machine</a>)</p> Answer 13: <p>(1) The older and more widely known applications are to regularity and solvability of PDEs of any order that are <strong>not</strong> necessarily elliptic. </p> <p>(2) The phenomenon of the <em>propagation of singularities</em> along certain submanifolds of the cotangent space is also another striking application. </p> <p>These refined notions of <em>hypoellipticity</em> and <em>solvability</em> as well as the propagation of singularities involve many fantastic results which are possible because of microlocal analysis. Even the statements of many of these results <strong>require</strong> microlocal analysis. </p> <p>(3) There are also several applications to several complex variables and a very active subfield called CR (for Cauchy Riemann) manifolds. </p> <p>Some of the standard references on these topics are the two volumes of F. Treves on <em>Introduction to Pseudodifferential and Fourier Integral Operators</em> and the L. Hormander's treatise (Volumes I, II, III, etc).</p> Answer 14: <p>Search for papers by Maarten de Hoop (Colorado School of Mines, then Purdue, now a Simons chair at Rice University), who has used microlocal analysis very extensively to study problems in global and exploration seismology and imaging. </p> Answer 15: <p>Broadly speaking, microlocal analysis helps one in 'geometrization' of certain results on the singularities of distributions. In this direction, one striking application of microlocal analysis is the Hormander condition for the existence of product of distributions. Study of nonlinear pdes and the <a href="http://projecteuclid.org/download/pdf_1/euclid.cmp/1104287114" rel="nofollow noreferrer">renormalization problem in Quantum Field Theory</a> are two major applications of product of distributions.</p> <p>An important tool in this context is the wavefront set of a distribution. Roughly speaking, the wavefront set locates the singularities of a distribution in space and also helps in the analysis of the 'direction' of singularities. For <span class="math-container">$u\in \mathcal{D}'(\mathbb{R}^n)$</span> the wavefront set <span class="math-container">$WF(u)\subset T^{*}(\mathbb{R}^n)$</span>. If <span class="math-container">$(x,\xi)\in WF(u)$</span> then <span class="math-container">$\xi\in \mathbb{R}^{n}-\{0\}$</span> is the direction of 'propagation of singularity' that is located at <span class="math-container">$x\in\mathbb{R}^{n}$</span>.</p> <p><strong>Hormander condition for existence of product:</strong> Let <span class="math-container">$u,v\in \mathcal{D}'({\mathbb{R}^n})$</span>. If for every <span class="math-container">$(x,\xi)\in WF(u)$</span> we have <span class="math-container">$(x,-\xi)\notin WF(v)$</span> then the product <span class="math-container">$uv$</span> exists and the Leibniz rule <span class="math-container">$\partial(uv)=\partial(u)v+u\partial(v)$</span> holds. Moreover, <span class="math-container">$WF(uv)\subseteq WF(u)\cup WF(v)\cup \{(x,\xi+\eta):(x,\xi)\in WF(u), (x,\eta)\in WF(v)\}$</span>.</p> <p>We can note the following from the above condition:</p> <ol> <li>If the point of singularity is same for both distributions then for the product to exist the 'directions' of the singularity <em>should not cancel</em> each other.</li> <li>We can also estimate the wavefront set of the product.</li> <li>The condition is only a sufficiency condition. The condition only provides information on the product of distributions that are relevant for pdes - those that satisfy the Lebiniz rule. For example, the Heaviside function can be multiplied with itself but the Lebniz rule does not hold for the square of Heaviside function. So, the Hormander condition rules out square of Heaviside function.</li> </ol> Answer 16: <p>A pretty recent and impressive application of microlocal analysis is to analytic number theory or automorphic forms, via the <a href="https://arxiv.org/pdf/1805.07750.pdf" rel="noreferrer">orbit method as developed by Nelson and Venkatesh</a>. In <em>loc. cit.</em> they set the stage for the theory specifically thought to be applied to problems in the area of automorphic forms (trace formulas, moments, subconvexity, and more widely problems on large families of automorphic forms).</p> <p>A brilliant outcome of this setting is the proof of <a href="https://arxiv.org/pdf/2109.15230.pdf" rel="noreferrer">subconvexity for all L-functions</a> of general linear groups, by Nelson.</p> Answer 17: <p>Microlocal analysis is used in <strong>computed tomography</strong> and other tomographic <strong>imaging techniques</strong> e.g. in <strong>medicine</strong> .</p> <p>Specifically, it is used to describe which wavefront sets (here: boundaries of objects, e.g. of organs in the human body) can or can not be detected by a specific tomographic measurement setup and also helps to understand reconstruction artifacts and develop strategies to overcome these. Details can be found for example in</p> <ul> <li>the overview article <a href="https://sites.tufts.edu/tquinto/files/2021/01//Wavefront-main.pdf" rel="nofollow noreferrer">Microlocal analysis in tomography</a> by Krishnan and Quinto and</li> <li>the &quot;Mathematics of Planet Earth&quot; article <a href="http://mpe.dimacs.rutgers.edu/2013/09/05/microlocal-analysis-and-imaging/" rel="nofollow noreferrer">Microlocal Analysis and Imaging</a>.</li> </ul> Answer 18: <p>Here are some examples:</p> <ol> <li>Helgason conjecture: Generically, the spherical principal series representations of a connected real semisimple Lie group $G$ are isomorphic via the Poisson transform to translation representations of $G$ on the spaces of joint eigenfunctions on the symmetric space $X=G/K$. The conjecture was proved in Kashiwara et al., Ann. Math. 107 (1978) using the boundary value theory in Kashiwara, Oshima, Ann. Math. 106 (1977). The inverse of the Poisson transform is shown to be a boundary value map defined by transforming, microlocally at the conormal bundle of a corner (Weyl chamber at infinity), the differential operators of the eigensystem on $X$ to simple normal form.</li> <li>Second term in Weyl asymptotics: The number $N(\lambda)$ of eigenvalues $\leq\lambda$ for the Dirichlet problem in $\Omega\subset\mathbb{R}^3$ of the Laplacian has the asymptotics $N(\lambda)=c_1\lambda^{3/2}+c_2\lambda+o(\lambda)$ if the closed geodesics cover only a set of measure zero in the unit sphere bundle. Up to constant factors $c_1$ and $c_2$ are the volumes of $\Omega$ and of its boundary $\partial\Omega$. This was proved by Ivrii (1980); see Corollary 29.3.4 in volume 4 of Hörmander's treatise. The existence of a two-term asymptotics under certain assumptions had been a long-standing conjecture.</li> <li>Distribution of scattering poles: The Lax--Phillips scattering matrix for acoustic scattering by a non-trapping obstacle has no poles in a logarithmic neighbourhood of the real axis. This was conjectured by Lax and Phillips in their 1967 book (Conjecture (a) on page 155) and proved by Melrose and Sjöstrand, CPAM 35 (1982), as a corollary of their results on propagation of wavefront sets in boundary problems.</li> <li>Controllability in control theory: The solutions of a wave equation are controllable in time $T&gt;0$ from a subset $\Gamma$ in the space-time boundary iff every geometric optics ray of time length $T$ intersects $\Gamma$. This was shown in Bardos, Lebeau, Rauch, SIAM J. Control Opt. 30 (1992); the proof uses the results of Melrose and Sjöstrand mentioned above.</li> </ol> <p>The list can be continued with trace formulae, applications to inverse problems, and more. Many results have been developed further in semiclassical analysis, a modern variant of microlocal analysis, which works under weaker assumptions, and is applicable to Schrödinger's equation, for example.</p> Answer 19: <p>There are striking applications in dynamical systems, due to Dyatlov and Zworski, where dynamical zeta functions were analysed using techniques from microlocal analysis. These zeta functions have a long history in dynamical systems and microlocal analysis is giving a range of new tools. </p> Answer 20: <p>Although microlocal analysis was developed originally exclusively for linear problems, it has played an increasingly important role in nonlinear PDE via what's known as paradifferential calculus. There are some nice write-ups about this. Here are a few:</p> <p><a href="https://www.math.ucla.edu/%7Etao/247b.1.07w/notes6.pdf" rel="nofollow noreferrer">https://www.math.ucla.edu/~tao/247b.1.07w/notes6.pdf</a></p> <p><a href="http://www.ams.org/notices/201007/rtx100700858p.pdf" rel="nofollow noreferrer">http://www.ams.org/notices/201007/rtx100700858p.pdf</a></p> <p><a href="http://www.unc.edu/math/Faculty/met/micro2.pdf" rel="nofollow noreferrer">http://www.unc.edu/math/Faculty/met/micro2.pdf</a> (<a href="https://web.archive.org/web/20180114042916/http://www.unc.edu/math/Faculty/met/micro2.pdf" rel="nofollow noreferrer">Wayback Machine</a>)</p> Answer 21: <p>(1) The older and more widely known applications are to regularity and solvability of PDEs of any order that are <strong>not</strong> necessarily elliptic. </p> <p>(2) The phenomenon of the <em>propagation of singularities</em> along certain submanifolds of the cotangent space is also another striking application. </p> <p>These refined notions of <em>hypoellipticity</em> and <em>solvability</em> as well as the propagation of singularities involve many fantastic results which are possible because of microlocal analysis. Even the statements of many of these results <strong>require</strong> microlocal analysis. </p> <p>(3) There are also several applications to several complex variables and a very active subfield called CR (for Cauchy Riemann) manifolds. </p> <p>Some of the standard references on these topics are the two volumes of F. Treves on <em>Introduction to Pseudodifferential and Fourier Integral Operators</em> and the L. Hormander's treatise (Volumes I, II, III, etc).</p> Answer 22: <p>Search for papers by Maarten de Hoop (Colorado School of Mines, then Purdue, now a Simons chair at Rice University), who has used microlocal analysis very extensively to study problems in global and exploration seismology and imaging. </p> Answer 23: <p>Broadly speaking, microlocal analysis helps one in 'geometrization' of certain results on the singularities of distributions. In this direction, one striking application of microlocal analysis is the Hormander condition for the existence of product of distributions. Study of nonlinear pdes and the <a href="http://projecteuclid.org/download/pdf_1/euclid.cmp/1104287114" rel="nofollow noreferrer">renormalization problem in Quantum Field Theory</a> are two major applications of product of distributions.</p> <p>An important tool in this context is the wavefront set of a distribution. Roughly speaking, the wavefront set locates the singularities of a distribution in space and also helps in the analysis of the 'direction' of singularities. For <span class="math-container">$u\in \mathcal{D}'(\mathbb{R}^n)$</span> the wavefront set <span class="math-container">$WF(u)\subset T^{*}(\mathbb{R}^n)$</span>. If <span class="math-container">$(x,\xi)\in WF(u)$</span> then <span class="math-container">$\xi\in \mathbb{R}^{n}-\{0\}$</span> is the direction of 'propagation of singularity' that is located at <span class="math-container">$x\in\mathbb{R}^{n}$</span>.</p> <p><strong>Hormander condition for existence of product:</strong> Let <span class="math-container">$u,v\in \mathcal{D}'({\mathbb{R}^n})$</span>. If for every <span class="math-container">$(x,\xi)\in WF(u)$</span> we have <span class="math-container">$(x,-\xi)\notin WF(v)$</span> then the product <span class="math-container">$uv$</span> exists and the Leibniz rule <span class="math-container">$\partial(uv)=\partial(u)v+u\partial(v)$</span> holds. Moreover, <span class="math-container">$WF(uv)\subseteq WF(u)\cup WF(v)\cup \{(x,\xi+\eta):(x,\xi)\in WF(u), (x,\eta)\in WF(v)\}$</span>.</p> <p>We can note the following from the above condition:</p> <ol> <li>If the point of singularity is same for both distributions then for the product to exist the 'directions' of the singularity <em>should not cancel</em> each other.</li> <li>We can also estimate the wavefront set of the product.</li> <li>The condition is only a sufficiency condition. The condition only provides information on the product of distributions that are relevant for pdes - those that satisfy the Lebiniz rule. For example, the Heaviside function can be multiplied with itself but the Lebniz rule does not hold for the square of Heaviside function. So, the Hormander condition rules out square of Heaviside function.</li> </ol> Answer 24: <p>A pretty recent and impressive application of microlocal analysis is to analytic number theory or automorphic forms, via the <a href="https://arxiv.org/pdf/1805.07750.pdf" rel="noreferrer">orbit method as developed by Nelson and Venkatesh</a>. In <em>loc. cit.</em> they set the stage for the theory specifically thought to be applied to problems in the area of automorphic forms (trace formulas, moments, subconvexity, and more widely problems on large families of automorphic forms).</p> <p>A brilliant outcome of this setting is the proof of <a href="https://arxiv.org/pdf/2109.15230.pdf" rel="noreferrer">subconvexity for all L-functions</a> of general linear groups, by Nelson.</p> Answer 25: <p>Microlocal analysis is used in <strong>computed tomography</strong> and other tomographic <strong>imaging techniques</strong> e.g. in <strong>medicine</strong> .</p> <p>Specifically, it is used to describe which wavefront sets (here: boundaries of objects, e.g. of organs in the human body) can or can not be detected by a specific tomographic measurement setup and also helps to understand reconstruction artifacts and develop strategies to overcome these. Details can be found for example in</p> <ul> <li>the overview article <a href="https://sites.tufts.edu/tquinto/files/2021/01//Wavefront-main.pdf" rel="nofollow noreferrer">Microlocal analysis in tomography</a> by Krishnan and Quinto and</li> <li>the &quot;Mathematics of Planet Earth&quot; article <a href="http://mpe.dimacs.rutgers.edu/2013/09/05/microlocal-analysis-and-imaging/" rel="nofollow noreferrer">Microlocal Analysis and Imaging</a>.</li> </ul> Answer 26: <p>Here are some examples:</p> <ol> <li>Helgason conjecture: Generically, the spherical principal series representations of a connected real semisimple Lie group $G$ are isomorphic via the Poisson transform to translation representations of $G$ on the spaces of joint eigenfunctions on the symmetric space $X=G/K$. The conjecture was proved in Kashiwara et al., Ann. Math. 107 (1978) using the boundary value theory in Kashiwara, Oshima, Ann. Math. 106 (1977). The inverse of the Poisson transform is shown to be a boundary value map defined by transforming, microlocally at the conormal bundle of a corner (Weyl chamber at infinity), the differential operators of the eigensystem on $X$ to simple normal form.</li> <li>Second term in Weyl asymptotics: The number $N(\lambda)$ of eigenvalues $\leq\lambda$ for the Dirichlet problem in $\Omega\subset\mathbb{R}^3$ of the Laplacian has the asymptotics $N(\lambda)=c_1\lambda^{3/2}+c_2\lambda+o(\lambda)$ if the closed geodesics cover only a set of measure zero in the unit sphere bundle. Up to constant factors $c_1$ and $c_2$ are the volumes of $\Omega$ and of its boundary $\partial\Omega$. This was proved by Ivrii (1980); see Corollary 29.3.4 in volume 4 of Hörmander's treatise. The existence of a two-term asymptotics under certain assumptions had been a long-standing conjecture.</li> <li>Distribution of scattering poles: The Lax--Phillips scattering matrix for acoustic scattering by a non-trapping obstacle has no poles in a logarithmic neighbourhood of the real axis. This was conjectured by Lax and Phillips in their 1967 book (Conjecture (a) on page 155) and proved by Melrose and Sjöstrand, CPAM 35 (1982), as a corollary of their results on propagation of wavefront sets in boundary problems.</li> <li>Controllability in control theory: The solutions of a wave equation are controllable in time $T&gt;0$ from a subset $\Gamma$ in the space-time boundary iff every geometric optics ray of time length $T$ intersects $\Gamma$. This was shown in Bardos, Lebeau, Rauch, SIAM J. Control Opt. 30 (1992); the proof uses the results of Melrose and Sjöstrand mentioned above.</li> </ol> <p>The list can be continued with trace formulae, applications to inverse problems, and more. Many results have been developed further in semiclassical analysis, a modern variant of microlocal analysis, which works under weaker assumptions, and is applicable to Schrödinger's equation, for example.</p> Answer 27: <p>There are striking applications in dynamical systems, due to Dyatlov and Zworski, where dynamical zeta functions were analysed using techniques from microlocal analysis. These zeta functions have a long history in dynamical systems and microlocal analysis is giving a range of new tools. </p> Answer 28: <p>Although microlocal analysis was developed originally exclusively for linear problems, it has played an increasingly important role in nonlinear PDE via what's known as paradifferential calculus. There are some nice write-ups about this. Here are a few:</p> <p><a href="https://www.math.ucla.edu/%7Etao/247b.1.07w/notes6.pdf" rel="nofollow noreferrer">https://www.math.ucla.edu/~tao/247b.1.07w/notes6.pdf</a></p> <p><a href="http://www.ams.org/notices/201007/rtx100700858p.pdf" rel="nofollow noreferrer">http://www.ams.org/notices/201007/rtx100700858p.pdf</a></p> <p><a href="http://www.unc.edu/math/Faculty/met/micro2.pdf" rel="nofollow noreferrer">http://www.unc.edu/math/Faculty/met/micro2.pdf</a> (<a href="https://web.archive.org/web/20180114042916/http://www.unc.edu/math/Faculty/met/micro2.pdf" rel="nofollow noreferrer">Wayback Machine</a>)</p> Answer 29: <p>(1) The older and more widely known applications are to regularity and solvability of PDEs of any order that are <strong>not</strong> necessarily elliptic. </p> <p>(2) The phenomenon of the <em>propagation of singularities</em> along certain submanifolds of the cotangent space is also another striking application. </p> <p>These refined notions of <em>hypoellipticity</em> and <em>solvability</em> as well as the propagation of singularities involve many fantastic results which are possible because of microlocal analysis. Even the statements of many of these results <strong>require</strong> microlocal analysis. </p> <p>(3) There are also several applications to several complex variables and a very active subfield called CR (for Cauchy Riemann) manifolds. </p> <p>Some of the standard references on these topics are the two volumes of F. Treves on <em>Introduction to Pseudodifferential and Fourier Integral Operators</em> and the L. Hormander's treatise (Volumes I, II, III, etc).</p> Answer 30: <p>Search for papers by Maarten de Hoop (Colorado School of Mines, then Purdue, now a Simons chair at Rice University), who has used microlocal analysis very extensively to study problems in global and exploration seismology and imaging. </p> Answer 31: <p>Broadly speaking, microlocal analysis helps one in 'geometrization' of certain results on the singularities of distributions. In this direction, one striking application of microlocal analysis is the Hormander condition for the existence of product of distributions. Study of nonlinear pdes and the <a href="http://projecteuclid.org/download/pdf_1/euclid.cmp/1104287114" rel="nofollow noreferrer">renormalization problem in Quantum Field Theory</a> are two major applications of product of distributions.</p> <p>An important tool in this context is the wavefront set of a distribution. Roughly speaking, the wavefront set locates the singularities of a distribution in space and also helps in the analysis of the 'direction' of singularities. For <span class="math-container">$u\in \mathcal{D}'(\mathbb{R}^n)$</span> the wavefront set <span class="math-container">$WF(u)\subset T^{*}(\mathbb{R}^n)$</span>. If <span class="math-container">$(x,\xi)\in WF(u)$</span> then <span class="math-container">$\xi\in \mathbb{R}^{n}-\{0\}$</span> is the direction of 'propagation of singularity' that is located at <span class="math-container">$x\in\mathbb{R}^{n}$</span>.</p> <p><strong>Hormander condition for existence of product:</strong> Let <span class="math-container">$u,v\in \mathcal{D}'({\mathbb{R}^n})$</span>. If for every <span class="math-container">$(x,\xi)\in WF(u)$</span> we have <span class="math-container">$(x,-\xi)\notin WF(v)$</span> then the product <span class="math-container">$uv$</span> exists and the Leibniz rule <span class="math-container">$\partial(uv)=\partial(u)v+u\partial(v)$</span> holds. Moreover, <span class="math-container">$WF(uv)\subseteq WF(u)\cup WF(v)\cup \{(x,\xi+\eta):(x,\xi)\in WF(u), (x,\eta)\in WF(v)\}$</span>.</p> <p>We can note the following from the above condition:</p> <ol> <li>If the point of singularity is same for both distributions then for the product to exist the 'directions' of the singularity <em>should not cancel</em> each other.</li> <li>We can also estimate the wavefront set of the product.</li> <li>The condition is only a sufficiency condition. The condition only provides information on the product of distributions that are relevant for pdes - those that satisfy the Lebiniz rule. For example, the Heaviside function can be multiplied with itself but the Lebniz rule does not hold for the square of Heaviside function. So, the Hormander condition rules out square of Heaviside function.</li> </ol> Answer 32: <p>A pretty recent and impressive application of microlocal analysis is to analytic number theory or automorphic forms, via the <a href="https://arxiv.org/pdf/1805.07750.pdf" rel="noreferrer">orbit method as developed by Nelson and Venkatesh</a>. In <em>loc. cit.</em> they set the stage for the theory specifically thought to be applied to problems in the area of automorphic forms (trace formulas, moments, subconvexity, and more widely problems on large families of automorphic forms).</p> <p>A brilliant outcome of this setting is the proof of <a href="https://arxiv.org/pdf/2109.15230.pdf" rel="noreferrer">subconvexity for all L-functions</a> of general linear groups, by Nelson.</p>
Is $x^{2k+1} - 7x^2 + 1$ irreducible?
258,914
28
https://mathoverflow.net/questions/258914/is-x2k1-7x2-1-irreducible
<blockquote> <p><strong>Question.</strong> Is the polynomial $x^{2k+1} - 7x^2 + 1$ irreducible over $\mathbb{Q}$ for every positive integer $k$?</p> </blockquote> <p>It is irreducible for all positive integers $k \leq 800$.</p>
[ { "answer_id": 259001, "score": 43, "text": "<p>Here is a proof, based on a trick that can be used to prove that\n$x^n + x + 1$ is irreducible when $n \\not\\equiv 2 \\bmod 3$.</p>\n\n<p>We work with Laurent polynomials in $R = \\mathbb Z[x,x^{-1}]$; note that\n$R$ has unit group $R^\\times = \\pm x^{\\mathbb Z}$.\nWe observe that for $f \\in R$, the sum of the squares of the coefficients\nis given by\n$$\\|f\\| = \\int_0^1 |f(e^{2 \\pi i t})|^2\\,dt\n = \\int_0^1 f(e^{2 \\pi i t}) f(e^{-2 \\pi i t})\\,dt\n = \\int_0^1 f(x) f(x^{-1})\\big|_{x = e^{2 \\pi i t}}\\,dt .$$\nNow assume that $f(x) = g(x) h(x)$. Then, since\n$f(x) f(x^{-1}) = \\bigl(g(x)h(x^{-1})\\bigr)\\bigl(g(x^{-1})h(x)\\bigr)$,\n$G(x) := g(x) h(x^{-1})$ satisfies $\\|G\\| = \\|f\\|$\nand $G(x) G(x^{-1}) = f(x) f(x^{-1})$.</p>\n\n<p>Now we consider $f(x) = x^n - 7 x^2 + 1$; then $\\|f\\| = 51$.\nIf $f = g h$ as above, then write\n$G(x) = \\pm x^m(1 + a_1 x + a_2 x^2 + \\ldots)$\nand $G(x^{-1}) = \\pm x^l(1 + b_1 x + b_2 x^2 + \\ldots)$.\nThe relation $G(x) G(x^{-1}) = f(x) f(x^{-1})$ translates into\n(equality of signs and)\n$$(1 + a_1 x + \\ldots)(1 + b_1 x + \\ldots) = 1 - 7 x^2 + O(x^{n-2}).$$\nAssuming that $n &gt; 40$ and considering terms up to $x^{20}$, one can\ncheck (see below) that the only solution such that\n$a_1^2 + a_2^2 + \\ldots + a_{20}^2 + b_1^2 + b_2^2 + \\ldots + b_{20}^2\\le 49$\nis, up to the substitution $x \\leftarrow x^{-1}$, given by\n$1 + a_1 x + \\ldots = 1 - 7x^2 + O(x^{21})$,\n$1 + b_1 x + \\ldots = 1 + O(x^{21})$. Since the $-7$ (together with the\nleading and trailing 1) exhausts our allowance\nfor the sum of squares of the coefficients, all other coefficients\nmust be zero, and we obtain that $G(x) = \\pm x^a f(x)$\nor $G(x) = \\pm x^a f(x^{-1})$. Modulo interchanging $g$ and $h$,\nwe can assume that $g(x) h(x^{-1}) = \\pm x^a f(x)$,\nso $h(x^{-1}) = \\pm x^a f(x)/g(x) = \\pm x^a h(x)$,\nand $x^{\\deg h} h(x^{-1})$ divides $f(x)$.\nThis implies that $h(x)$ divides $x^n f(x^{-1})$, so $h(x)$ must divide\n$$f(x) - x^n f(x^{-1}) = 7 x^2 (x^{n-4} - 1).$$\nSo $h$ also divides\n$$f(x) - x^4 (x^{n-4} - 1) = x^4 - 7 x^2 + 1 = (x^2-3x+1)(x^2+3x+1).$$\nSince $h$ also divides $f$, it must divide the difference $x^n - x^4$,\nwhich for $n \\neq 4$ it clearly doesn't, since the quartic has no\nroots of absolute value 0 or 1; contradiction.</p>\n\n<p>The argument shows that $x^n - 7 x^2 + 1$ is irreducible for $n &gt; 40$;\nfor smaller $n$, we can ask the Computer Algebra System we trust.\nThis gives:</p>\n\n<p><strong>Theorem.</strong> $x^n - 7 x^2 + 1$ is irreducible over $\\mathbb Q$\nfor all positive integers $n$ except $n=4$.</p>\n\n<p><strong>ADDED 2017-01-08:</strong>\nAfter re-checking the computations, I realized that there\nwas a small mistake that ruled out some partial solutions prematurely.\nIt looks like one needs to consider terms up to $x^{20}$.\n<a href=\"http://www.mathe2.uni-bayreuth.de/stoll/MO/MO-258914.magma\">Here</a> is a file with MAGMA code that verifies that\nthere are no other solutions.</p>\n" }, { "answer_id": 258922, "score": 15, "text": "<p>Not an answer, but an observation, which, if proved, will probably give an answer.</p>\n\n<p>For all $k$ that I have computed, your polynomial has two roots smaller than $1$ in absolute value (for large $k$ these converge to $\\pm\\frac{1}{\\sqrt{7}}.$), and the rest of the roots are larger than $1$ in absolute value. Let's assume this is true. Since the constant term of a factor has to have absolute value $1$ each factor should have one of these roots, so there are at most $2$ irreducible factors.</p>\n" }, { "answer_id": 258923, "score": 10, "text": "<p>EDIT: I'm afraid that this answer is wrong and I hope you have some way of \"unaccepting\" it. The reason is that when taking conjugates of $\\frac{1}{\\sqrt{7}-\\beta_1}$ we're allowed to freely change the sign of $\\sqrt{7}$ (unless $\\sqrt{7}$ is contained in the field generated by $\\beta_1$, which seems rather unlikely). And $\\frac{1}{-\\sqrt{7}-\\beta_1}$ is greater than $1$ in absolute value. I found this mistake only when I noticed that my \"proof\" gives that $x^4-7x^2+1=(x^2+3x+1)(x^2-3x+1)$ is irreducible. It should work with an integer square in place of $7$. Maybe someone can make it work also for $7$? I apologize for raising false hopes!</p>\n\n<p>The question is equivalent (by replacing $x$ with $\\frac{1}{x}$) to asking about the irreducibility of $A(x) = x^{2k+1}-7x^{2k-1}+1 = x^{2k-1}(x^2-7)+1$ over $\\mathbb{Q}$. An application of Rouché's theorem shows that $A(x)$ has exactly $2k-1$ zeroes (counted with multiplicity) of absolute value less than $1$. There remain two real zeroes $\\beta_1 &lt; -1$ and $\\beta_2 &gt; 1$. One can deduce from the equation they satisfy that $|\\beta_i - (-1)^i\\sqrt{7}| \\leq \\frac{1}{\\sqrt{7}+1}$ ($i=1,2$).</p>\n\n<p>Now suppose that $A$ factorizes as $A = BC$ with monic integer polynomials $B, C$. Wlog $B(\\beta_1) = 0$. If $B(\\beta_2) \\neq 0$, then\n$$\\beta_1^{2k-1}(\\beta_1+\\sqrt{7}) = \\frac{1}{\\sqrt{7}-\\beta_1}$$\nis an algebraic integer, all conjugates of which are less than $1$ in absolute value (the right hand side is less than $\\frac{1}{2\\sqrt{7}-\\frac{1}{\\sqrt{7}+1}}&lt;1$ and all its conjugates are of absolute value at most $\\frac{1}{\\sqrt{7}-1}&lt;1$).</p>\n\n<p>So it would have to be $0$, which is absurd. So $B(\\beta_2)=0$. But now all roots of $C$ are less than $1$ in absolute value, while its constant term is $1$. It follows that $C=1$ is constant and so $A$ is irreducible.</p>\n" }, { "answer_id": 259001, "score": 43, "text": "<p>Here is a proof, based on a trick that can be used to prove that\n$x^n + x + 1$ is irreducible when $n \\not\\equiv 2 \\bmod 3$.</p>\n\n<p>We work with Laurent polynomials in $R = \\mathbb Z[x,x^{-1}]$; note that\n$R$ has unit group $R^\\times = \\pm x^{\\mathbb Z}$.\nWe observe that for $f \\in R$, the sum of the squares of the coefficients\nis given by\n$$\\|f\\| = \\int_0^1 |f(e^{2 \\pi i t})|^2\\,dt\n = \\int_0^1 f(e^{2 \\pi i t}) f(e^{-2 \\pi i t})\\,dt\n = \\int_0^1 f(x) f(x^{-1})\\big|_{x = e^{2 \\pi i t}}\\,dt .$$\nNow assume that $f(x) = g(x) h(x)$. Then, since\n$f(x) f(x^{-1}) = \\bigl(g(x)h(x^{-1})\\bigr)\\bigl(g(x^{-1})h(x)\\bigr)$,\n$G(x) := g(x) h(x^{-1})$ satisfies $\\|G\\| = \\|f\\|$\nand $G(x) G(x^{-1}) = f(x) f(x^{-1})$.</p>\n\n<p>Now we consider $f(x) = x^n - 7 x^2 + 1$; then $\\|f\\| = 51$.\nIf $f = g h$ as above, then write\n$G(x) = \\pm x^m(1 + a_1 x + a_2 x^2 + \\ldots)$\nand $G(x^{-1}) = \\pm x^l(1 + b_1 x + b_2 x^2 + \\ldots)$.\nThe relation $G(x) G(x^{-1}) = f(x) f(x^{-1})$ translates into\n(equality of signs and)\n$$(1 + a_1 x + \\ldots)(1 + b_1 x + \\ldots) = 1 - 7 x^2 + O(x^{n-2}).$$\nAssuming that $n &gt; 40$ and considering terms up to $x^{20}$, one can\ncheck (see below) that the only solution such that\n$a_1^2 + a_2^2 + \\ldots + a_{20}^2 + b_1^2 + b_2^2 + \\ldots + b_{20}^2\\le 49$\nis, up to the substitution $x \\leftarrow x^{-1}$, given by\n$1 + a_1 x + \\ldots = 1 - 7x^2 + O(x^{21})$,\n$1 + b_1 x + \\ldots = 1 + O(x^{21})$. Since the $-7$ (together with the\nleading and trailing 1) exhausts our allowance\nfor the sum of squares of the coefficients, all other coefficients\nmust be zero, and we obtain that $G(x) = \\pm x^a f(x)$\nor $G(x) = \\pm x^a f(x^{-1})$. Modulo interchanging $g$ and $h$,\nwe can assume that $g(x) h(x^{-1}) = \\pm x^a f(x)$,\nso $h(x^{-1}) = \\pm x^a f(x)/g(x) = \\pm x^a h(x)$,\nand $x^{\\deg h} h(x^{-1})$ divides $f(x)$.\nThis implies that $h(x)$ divides $x^n f(x^{-1})$, so $h(x)$ must divide\n$$f(x) - x^n f(x^{-1}) = 7 x^2 (x^{n-4} - 1).$$\nSo $h$ also divides\n$$f(x) - x^4 (x^{n-4} - 1) = x^4 - 7 x^2 + 1 = (x^2-3x+1)(x^2+3x+1).$$\nSince $h$ also divides $f$, it must divide the difference $x^n - x^4$,\nwhich for $n \\neq 4$ it clearly doesn't, since the quartic has no\nroots of absolute value 0 or 1; contradiction.</p>\n\n<p>The argument shows that $x^n - 7 x^2 + 1$ is irreducible for $n &gt; 40$;\nfor smaller $n$, we can ask the Computer Algebra System we trust.\nThis gives:</p>\n\n<p><strong>Theorem.</strong> $x^n - 7 x^2 + 1$ is irreducible over $\\mathbb Q$\nfor all positive integers $n$ except $n=4$.</p>\n\n<p><strong>ADDED 2017-01-08:</strong>\nAfter re-checking the computations, I realized that there\nwas a small mistake that ruled out some partial solutions prematurely.\nIt looks like one needs to consider terms up to $x^{20}$.\n<a href=\"http://www.mathe2.uni-bayreuth.de/stoll/MO/MO-258914.magma\">Here</a> is a file with MAGMA code that verifies that\nthere are no other solutions.</p>\n" }, { "answer_id": 258922, "score": 15, "text": "<p>Not an answer, but an observation, which, if proved, will probably give an answer.</p>\n\n<p>For all $k$ that I have computed, your polynomial has two roots smaller than $1$ in absolute value (for large $k$ these converge to $\\pm\\frac{1}{\\sqrt{7}}.$), and the rest of the roots are larger than $1$ in absolute value. Let's assume this is true. Since the constant term of a factor has to have absolute value $1$ each factor should have one of these roots, so there are at most $2$ irreducible factors.</p>\n" }, { "answer_id": 258923, "score": 10, "text": "<p>EDIT: I'm afraid that this answer is wrong and I hope you have some way of \"unaccepting\" it. The reason is that when taking conjugates of $\\frac{1}{\\sqrt{7}-\\beta_1}$ we're allowed to freely change the sign of $\\sqrt{7}$ (unless $\\sqrt{7}$ is contained in the field generated by $\\beta_1$, which seems rather unlikely). And $\\frac{1}{-\\sqrt{7}-\\beta_1}$ is greater than $1$ in absolute value. I found this mistake only when I noticed that my \"proof\" gives that $x^4-7x^2+1=(x^2+3x+1)(x^2-3x+1)$ is irreducible. It should work with an integer square in place of $7$. Maybe someone can make it work also for $7$? I apologize for raising false hopes!</p>\n\n<p>The question is equivalent (by replacing $x$ with $\\frac{1}{x}$) to asking about the irreducibility of $A(x) = x^{2k+1}-7x^{2k-1}+1 = x^{2k-1}(x^2-7)+1$ over $\\mathbb{Q}$. An application of Rouché's theorem shows that $A(x)$ has exactly $2k-1$ zeroes (counted with multiplicity) of absolute value less than $1$. There remain two real zeroes $\\beta_1 &lt; -1$ and $\\beta_2 &gt; 1$. One can deduce from the equation they satisfy that $|\\beta_i - (-1)^i\\sqrt{7}| \\leq \\frac{1}{\\sqrt{7}+1}$ ($i=1,2$).</p>\n\n<p>Now suppose that $A$ factorizes as $A = BC$ with monic integer polynomials $B, C$. Wlog $B(\\beta_1) = 0$. If $B(\\beta_2) \\neq 0$, then\n$$\\beta_1^{2k-1}(\\beta_1+\\sqrt{7}) = \\frac{1}{\\sqrt{7}-\\beta_1}$$\nis an algebraic integer, all conjugates of which are less than $1$ in absolute value (the right hand side is less than $\\frac{1}{2\\sqrt{7}-\\frac{1}{\\sqrt{7}+1}}&lt;1$ and all its conjugates are of absolute value at most $\\frac{1}{\\sqrt{7}-1}&lt;1$).</p>\n\n<p>So it would have to be $0$, which is absurd. So $B(\\beta_2)=0$. But now all roots of $C$ are less than $1$ in absolute value, while its constant term is $1$. It follows that $C=1$ is constant and so $A$ is irreducible.</p>\n" }, { "answer_id": 259001, "score": 43, "text": "<p>Here is a proof, based on a trick that can be used to prove that\n$x^n + x + 1$ is irreducible when $n \\not\\equiv 2 \\bmod 3$.</p>\n\n<p>We work with Laurent polynomials in $R = \\mathbb Z[x,x^{-1}]$; note that\n$R$ has unit group $R^\\times = \\pm x^{\\mathbb Z}$.\nWe observe that for $f \\in R$, the sum of the squares of the coefficients\nis given by\n$$\\|f\\| = \\int_0^1 |f(e^{2 \\pi i t})|^2\\,dt\n = \\int_0^1 f(e^{2 \\pi i t}) f(e^{-2 \\pi i t})\\,dt\n = \\int_0^1 f(x) f(x^{-1})\\big|_{x = e^{2 \\pi i t}}\\,dt .$$\nNow assume that $f(x) = g(x) h(x)$. Then, since\n$f(x) f(x^{-1}) = \\bigl(g(x)h(x^{-1})\\bigr)\\bigl(g(x^{-1})h(x)\\bigr)$,\n$G(x) := g(x) h(x^{-1})$ satisfies $\\|G\\| = \\|f\\|$\nand $G(x) G(x^{-1}) = f(x) f(x^{-1})$.</p>\n\n<p>Now we consider $f(x) = x^n - 7 x^2 + 1$; then $\\|f\\| = 51$.\nIf $f = g h$ as above, then write\n$G(x) = \\pm x^m(1 + a_1 x + a_2 x^2 + \\ldots)$\nand $G(x^{-1}) = \\pm x^l(1 + b_1 x + b_2 x^2 + \\ldots)$.\nThe relation $G(x) G(x^{-1}) = f(x) f(x^{-1})$ translates into\n(equality of signs and)\n$$(1 + a_1 x + \\ldots)(1 + b_1 x + \\ldots) = 1 - 7 x^2 + O(x^{n-2}).$$\nAssuming that $n &gt; 40$ and considering terms up to $x^{20}$, one can\ncheck (see below) that the only solution such that\n$a_1^2 + a_2^2 + \\ldots + a_{20}^2 + b_1^2 + b_2^2 + \\ldots + b_{20}^2\\le 49$\nis, up to the substitution $x \\leftarrow x^{-1}$, given by\n$1 + a_1 x + \\ldots = 1 - 7x^2 + O(x^{21})$,\n$1 + b_1 x + \\ldots = 1 + O(x^{21})$. Since the $-7$ (together with the\nleading and trailing 1) exhausts our allowance\nfor the sum of squares of the coefficients, all other coefficients\nmust be zero, and we obtain that $G(x) = \\pm x^a f(x)$\nor $G(x) = \\pm x^a f(x^{-1})$. Modulo interchanging $g$ and $h$,\nwe can assume that $g(x) h(x^{-1}) = \\pm x^a f(x)$,\nso $h(x^{-1}) = \\pm x^a f(x)/g(x) = \\pm x^a h(x)$,\nand $x^{\\deg h} h(x^{-1})$ divides $f(x)$.\nThis implies that $h(x)$ divides $x^n f(x^{-1})$, so $h(x)$ must divide\n$$f(x) - x^n f(x^{-1}) = 7 x^2 (x^{n-4} - 1).$$\nSo $h$ also divides\n$$f(x) - x^4 (x^{n-4} - 1) = x^4 - 7 x^2 + 1 = (x^2-3x+1)(x^2+3x+1).$$\nSince $h$ also divides $f$, it must divide the difference $x^n - x^4$,\nwhich for $n \\neq 4$ it clearly doesn't, since the quartic has no\nroots of absolute value 0 or 1; contradiction.</p>\n\n<p>The argument shows that $x^n - 7 x^2 + 1$ is irreducible for $n &gt; 40$;\nfor smaller $n$, we can ask the Computer Algebra System we trust.\nThis gives:</p>\n\n<p><strong>Theorem.</strong> $x^n - 7 x^2 + 1$ is irreducible over $\\mathbb Q$\nfor all positive integers $n$ except $n=4$.</p>\n\n<p><strong>ADDED 2017-01-08:</strong>\nAfter re-checking the computations, I realized that there\nwas a small mistake that ruled out some partial solutions prematurely.\nIt looks like one needs to consider terms up to $x^{20}$.\n<a href=\"http://www.mathe2.uni-bayreuth.de/stoll/MO/MO-258914.magma\">Here</a> is a file with MAGMA code that verifies that\nthere are no other solutions.</p>\n" }, { "answer_id": 258922, "score": 15, "text": "<p>Not an answer, but an observation, which, if proved, will probably give an answer.</p>\n\n<p>For all $k$ that I have computed, your polynomial has two roots smaller than $1$ in absolute value (for large $k$ these converge to $\\pm\\frac{1}{\\sqrt{7}}.$), and the rest of the roots are larger than $1$ in absolute value. Let's assume this is true. Since the constant term of a factor has to have absolute value $1$ each factor should have one of these roots, so there are at most $2$ irreducible factors.</p>\n" }, { "answer_id": 258923, "score": 10, "text": "<p>EDIT: I'm afraid that this answer is wrong and I hope you have some way of \"unaccepting\" it. The reason is that when taking conjugates of $\\frac{1}{\\sqrt{7}-\\beta_1}$ we're allowed to freely change the sign of $\\sqrt{7}$ (unless $\\sqrt{7}$ is contained in the field generated by $\\beta_1$, which seems rather unlikely). And $\\frac{1}{-\\sqrt{7}-\\beta_1}$ is greater than $1$ in absolute value. I found this mistake only when I noticed that my \"proof\" gives that $x^4-7x^2+1=(x^2+3x+1)(x^2-3x+1)$ is irreducible. It should work with an integer square in place of $7$. Maybe someone can make it work also for $7$? I apologize for raising false hopes!</p>\n\n<p>The question is equivalent (by replacing $x$ with $\\frac{1}{x}$) to asking about the irreducibility of $A(x) = x^{2k+1}-7x^{2k-1}+1 = x^{2k-1}(x^2-7)+1$ over $\\mathbb{Q}$. An application of Rouché's theorem shows that $A(x)$ has exactly $2k-1$ zeroes (counted with multiplicity) of absolute value less than $1$. There remain two real zeroes $\\beta_1 &lt; -1$ and $\\beta_2 &gt; 1$. One can deduce from the equation they satisfy that $|\\beta_i - (-1)^i\\sqrt{7}| \\leq \\frac{1}{\\sqrt{7}+1}$ ($i=1,2$).</p>\n\n<p>Now suppose that $A$ factorizes as $A = BC$ with monic integer polynomials $B, C$. Wlog $B(\\beta_1) = 0$. If $B(\\beta_2) \\neq 0$, then\n$$\\beta_1^{2k-1}(\\beta_1+\\sqrt{7}) = \\frac{1}{\\sqrt{7}-\\beta_1}$$\nis an algebraic integer, all conjugates of which are less than $1$ in absolute value (the right hand side is less than $\\frac{1}{2\\sqrt{7}-\\frac{1}{\\sqrt{7}+1}}&lt;1$ and all its conjugates are of absolute value at most $\\frac{1}{\\sqrt{7}-1}&lt;1$).</p>\n\n<p>So it would have to be $0$, which is absurd. So $B(\\beta_2)=0$. But now all roots of $C$ are less than $1$ in absolute value, while its constant term is $1$. It follows that $C=1$ is constant and so $A$ is irreducible.</p>\n" }, { "answer_id": 259001, "score": 43, "text": "<p>Here is a proof, based on a trick that can be used to prove that\n$x^n + x + 1$ is irreducible when $n \\not\\equiv 2 \\bmod 3$.</p>\n\n<p>We work with Laurent polynomials in $R = \\mathbb Z[x,x^{-1}]$; note that\n$R$ has unit group $R^\\times = \\pm x^{\\mathbb Z}$.\nWe observe that for $f \\in R$, the sum of the squares of the coefficients\nis given by\n$$\\|f\\| = \\int_0^1 |f(e^{2 \\pi i t})|^2\\,dt\n = \\int_0^1 f(e^{2 \\pi i t}) f(e^{-2 \\pi i t})\\,dt\n = \\int_0^1 f(x) f(x^{-1})\\big|_{x = e^{2 \\pi i t}}\\,dt .$$\nNow assume that $f(x) = g(x) h(x)$. Then, since\n$f(x) f(x^{-1}) = \\bigl(g(x)h(x^{-1})\\bigr)\\bigl(g(x^{-1})h(x)\\bigr)$,\n$G(x) := g(x) h(x^{-1})$ satisfies $\\|G\\| = \\|f\\|$\nand $G(x) G(x^{-1}) = f(x) f(x^{-1})$.</p>\n\n<p>Now we consider $f(x) = x^n - 7 x^2 + 1$; then $\\|f\\| = 51$.\nIf $f = g h$ as above, then write\n$G(x) = \\pm x^m(1 + a_1 x + a_2 x^2 + \\ldots)$\nand $G(x^{-1}) = \\pm x^l(1 + b_1 x + b_2 x^2 + \\ldots)$.\nThe relation $G(x) G(x^{-1}) = f(x) f(x^{-1})$ translates into\n(equality of signs and)\n$$(1 + a_1 x + \\ldots)(1 + b_1 x + \\ldots) = 1 - 7 x^2 + O(x^{n-2}).$$\nAssuming that $n &gt; 40$ and considering terms up to $x^{20}$, one can\ncheck (see below) that the only solution such that\n$a_1^2 + a_2^2 + \\ldots + a_{20}^2 + b_1^2 + b_2^2 + \\ldots + b_{20}^2\\le 49$\nis, up to the substitution $x \\leftarrow x^{-1}$, given by\n$1 + a_1 x + \\ldots = 1 - 7x^2 + O(x^{21})$,\n$1 + b_1 x + \\ldots = 1 + O(x^{21})$. Since the $-7$ (together with the\nleading and trailing 1) exhausts our allowance\nfor the sum of squares of the coefficients, all other coefficients\nmust be zero, and we obtain that $G(x) = \\pm x^a f(x)$\nor $G(x) = \\pm x^a f(x^{-1})$. Modulo interchanging $g$ and $h$,\nwe can assume that $g(x) h(x^{-1}) = \\pm x^a f(x)$,\nso $h(x^{-1}) = \\pm x^a f(x)/g(x) = \\pm x^a h(x)$,\nand $x^{\\deg h} h(x^{-1})$ divides $f(x)$.\nThis implies that $h(x)$ divides $x^n f(x^{-1})$, so $h(x)$ must divide\n$$f(x) - x^n f(x^{-1}) = 7 x^2 (x^{n-4} - 1).$$\nSo $h$ also divides\n$$f(x) - x^4 (x^{n-4} - 1) = x^4 - 7 x^2 + 1 = (x^2-3x+1)(x^2+3x+1).$$\nSince $h$ also divides $f$, it must divide the difference $x^n - x^4$,\nwhich for $n \\neq 4$ it clearly doesn't, since the quartic has no\nroots of absolute value 0 or 1; contradiction.</p>\n\n<p>The argument shows that $x^n - 7 x^2 + 1$ is irreducible for $n &gt; 40$;\nfor smaller $n$, we can ask the Computer Algebra System we trust.\nThis gives:</p>\n\n<p><strong>Theorem.</strong> $x^n - 7 x^2 + 1$ is irreducible over $\\mathbb Q$\nfor all positive integers $n$ except $n=4$.</p>\n\n<p><strong>ADDED 2017-01-08:</strong>\nAfter re-checking the computations, I realized that there\nwas a small mistake that ruled out some partial solutions prematurely.\nIt looks like one needs to consider terms up to $x^{20}$.\n<a href=\"http://www.mathe2.uni-bayreuth.de/stoll/MO/MO-258914.magma\">Here</a> is a file with MAGMA code that verifies that\nthere are no other solutions.</p>\n" }, { "answer_id": 258922, "score": 15, "text": "<p>Not an answer, but an observation, which, if proved, will probably give an answer.</p>\n\n<p>For all $k$ that I have computed, your polynomial has two roots smaller than $1$ in absolute value (for large $k$ these converge to $\\pm\\frac{1}{\\sqrt{7}}.$), and the rest of the roots are larger than $1$ in absolute value. Let's assume this is true. Since the constant term of a factor has to have absolute value $1$ each factor should have one of these roots, so there are at most $2$ irreducible factors.</p>\n" }, { "answer_id": 258923, "score": 10, "text": "<p>EDIT: I'm afraid that this answer is wrong and I hope you have some way of \"unaccepting\" it. The reason is that when taking conjugates of $\\frac{1}{\\sqrt{7}-\\beta_1}$ we're allowed to freely change the sign of $\\sqrt{7}$ (unless $\\sqrt{7}$ is contained in the field generated by $\\beta_1$, which seems rather unlikely). And $\\frac{1}{-\\sqrt{7}-\\beta_1}$ is greater than $1$ in absolute value. I found this mistake only when I noticed that my \"proof\" gives that $x^4-7x^2+1=(x^2+3x+1)(x^2-3x+1)$ is irreducible. It should work with an integer square in place of $7$. Maybe someone can make it work also for $7$? I apologize for raising false hopes!</p>\n\n<p>The question is equivalent (by replacing $x$ with $\\frac{1}{x}$) to asking about the irreducibility of $A(x) = x^{2k+1}-7x^{2k-1}+1 = x^{2k-1}(x^2-7)+1$ over $\\mathbb{Q}$. An application of Rouché's theorem shows that $A(x)$ has exactly $2k-1$ zeroes (counted with multiplicity) of absolute value less than $1$. There remain two real zeroes $\\beta_1 &lt; -1$ and $\\beta_2 &gt; 1$. One can deduce from the equation they satisfy that $|\\beta_i - (-1)^i\\sqrt{7}| \\leq \\frac{1}{\\sqrt{7}+1}$ ($i=1,2$).</p>\n\n<p>Now suppose that $A$ factorizes as $A = BC$ with monic integer polynomials $B, C$. Wlog $B(\\beta_1) = 0$. If $B(\\beta_2) \\neq 0$, then\n$$\\beta_1^{2k-1}(\\beta_1+\\sqrt{7}) = \\frac{1}{\\sqrt{7}-\\beta_1}$$\nis an algebraic integer, all conjugates of which are less than $1$ in absolute value (the right hand side is less than $\\frac{1}{2\\sqrt{7}-\\frac{1}{\\sqrt{7}+1}}&lt;1$ and all its conjugates are of absolute value at most $\\frac{1}{\\sqrt{7}-1}&lt;1$).</p>\n\n<p>So it would have to be $0$, which is absurd. So $B(\\beta_2)=0$. But now all roots of $C$ are less than $1$ in absolute value, while its constant term is $1$. It follows that $C=1$ is constant and so $A$ is irreducible.</p>\n" } ]
Question: Is $x^{2k+1} - 7x^2 + 1$ irreducible? Answer 1: <p>Here is a proof, based on a trick that can be used to prove that $x^n + x + 1$ is irreducible when $n \not\equiv 2 \bmod 3$.</p> <p>We work with Laurent polynomials in $R = \mathbb Z[x,x^{-1}]$; note that $R$ has unit group $R^\times = \pm x^{\mathbb Z}$. We observe that for $f \in R$, the sum of the squares of the coefficients is given by $$\|f\| = \int_0^1 |f(e^{2 \pi i t})|^2\,dt = \int_0^1 f(e^{2 \pi i t}) f(e^{-2 \pi i t})\,dt = \int_0^1 f(x) f(x^{-1})\big|_{x = e^{2 \pi i t}}\,dt .$$ Now assume that $f(x) = g(x) h(x)$. Then, since $f(x) f(x^{-1}) = \bigl(g(x)h(x^{-1})\bigr)\bigl(g(x^{-1})h(x)\bigr)$, $G(x) := g(x) h(x^{-1})$ satisfies $\|G\| = \|f\|$ and $G(x) G(x^{-1}) = f(x) f(x^{-1})$.</p> <p>Now we consider $f(x) = x^n - 7 x^2 + 1$; then $\|f\| = 51$. If $f = g h$ as above, then write $G(x) = \pm x^m(1 + a_1 x + a_2 x^2 + \ldots)$ and $G(x^{-1}) = \pm x^l(1 + b_1 x + b_2 x^2 + \ldots)$. The relation $G(x) G(x^{-1}) = f(x) f(x^{-1})$ translates into (equality of signs and) $$(1 + a_1 x + \ldots)(1 + b_1 x + \ldots) = 1 - 7 x^2 + O(x^{n-2}).$$ Assuming that $n &gt; 40$ and considering terms up to $x^{20}$, one can check (see below) that the only solution such that $a_1^2 + a_2^2 + \ldots + a_{20}^2 + b_1^2 + b_2^2 + \ldots + b_{20}^2\le 49$ is, up to the substitution $x \leftarrow x^{-1}$, given by $1 + a_1 x + \ldots = 1 - 7x^2 + O(x^{21})$, $1 + b_1 x + \ldots = 1 + O(x^{21})$. Since the $-7$ (together with the leading and trailing 1) exhausts our allowance for the sum of squares of the coefficients, all other coefficients must be zero, and we obtain that $G(x) = \pm x^a f(x)$ or $G(x) = \pm x^a f(x^{-1})$. Modulo interchanging $g$ and $h$, we can assume that $g(x) h(x^{-1}) = \pm x^a f(x)$, so $h(x^{-1}) = \pm x^a f(x)/g(x) = \pm x^a h(x)$, and $x^{\deg h} h(x^{-1})$ divides $f(x)$. This implies that $h(x)$ divides $x^n f(x^{-1})$, so $h(x)$ must divide $$f(x) - x^n f(x^{-1}) = 7 x^2 (x^{n-4} - 1).$$ So $h$ also divides $$f(x) - x^4 (x^{n-4} - 1) = x^4 - 7 x^2 + 1 = (x^2-3x+1)(x^2+3x+1).$$ Since $h$ also divides $f$, it must divide the difference $x^n - x^4$, which for $n \neq 4$ it clearly doesn't, since the quartic has no roots of absolute value 0 or 1; contradiction.</p> <p>The argument shows that $x^n - 7 x^2 + 1$ is irreducible for $n &gt; 40$; for smaller $n$, we can ask the Computer Algebra System we trust. This gives:</p> <p><strong>Theorem.</strong> $x^n - 7 x^2 + 1$ is irreducible over $\mathbb Q$ for all positive integers $n$ except $n=4$.</p> <p><strong>ADDED 2017-01-08:</strong> After re-checking the computations, I realized that there was a small mistake that ruled out some partial solutions prematurely. It looks like one needs to consider terms up to $x^{20}$. <a href="http://www.mathe2.uni-bayreuth.de/stoll/MO/MO-258914.magma">Here</a> is a file with MAGMA code that verifies that there are no other solutions.</p> Answer 2: <p>Not an answer, but an observation, which, if proved, will probably give an answer.</p> <p>For all $k$ that I have computed, your polynomial has two roots smaller than $1$ in absolute value (for large $k$ these converge to $\pm\frac{1}{\sqrt{7}}.$), and the rest of the roots are larger than $1$ in absolute value. Let's assume this is true. Since the constant term of a factor has to have absolute value $1$ each factor should have one of these roots, so there are at most $2$ irreducible factors.</p> Answer 3: <p>EDIT: I'm afraid that this answer is wrong and I hope you have some way of "unaccepting" it. The reason is that when taking conjugates of $\frac{1}{\sqrt{7}-\beta_1}$ we're allowed to freely change the sign of $\sqrt{7}$ (unless $\sqrt{7}$ is contained in the field generated by $\beta_1$, which seems rather unlikely). And $\frac{1}{-\sqrt{7}-\beta_1}$ is greater than $1$ in absolute value. I found this mistake only when I noticed that my "proof" gives that $x^4-7x^2+1=(x^2+3x+1)(x^2-3x+1)$ is irreducible. It should work with an integer square in place of $7$. Maybe someone can make it work also for $7$? I apologize for raising false hopes!</p> <p>The question is equivalent (by replacing $x$ with $\frac{1}{x}$) to asking about the irreducibility of $A(x) = x^{2k+1}-7x^{2k-1}+1 = x^{2k-1}(x^2-7)+1$ over $\mathbb{Q}$. An application of Rouché's theorem shows that $A(x)$ has exactly $2k-1$ zeroes (counted with multiplicity) of absolute value less than $1$. There remain two real zeroes $\beta_1 &lt; -1$ and $\beta_2 &gt; 1$. One can deduce from the equation they satisfy that $|\beta_i - (-1)^i\sqrt{7}| \leq \frac{1}{\sqrt{7}+1}$ ($i=1,2$).</p> <p>Now suppose that $A$ factorizes as $A = BC$ with monic integer polynomials $B, C$. Wlog $B(\beta_1) = 0$. If $B(\beta_2) \neq 0$, then $$\beta_1^{2k-1}(\beta_1+\sqrt{7}) = \frac{1}{\sqrt{7}-\beta_1}$$ is an algebraic integer, all conjugates of which are less than $1$ in absolute value (the right hand side is less than $\frac{1}{2\sqrt{7}-\frac{1}{\sqrt{7}+1}}&lt;1$ and all its conjugates are of absolute value at most $\frac{1}{\sqrt{7}-1}&lt;1$).</p> <p>So it would have to be $0$, which is absurd. So $B(\beta_2)=0$. But now all roots of $C$ are less than $1$ in absolute value, while its constant term is $1$. It follows that $C=1$ is constant and so $A$ is irreducible.</p> Answer 4: <p>Here is a proof, based on a trick that can be used to prove that $x^n + x + 1$ is irreducible when $n \not\equiv 2 \bmod 3$.</p> <p>We work with Laurent polynomials in $R = \mathbb Z[x,x^{-1}]$; note that $R$ has unit group $R^\times = \pm x^{\mathbb Z}$. We observe that for $f \in R$, the sum of the squares of the coefficients is given by $$\|f\| = \int_0^1 |f(e^{2 \pi i t})|^2\,dt = \int_0^1 f(e^{2 \pi i t}) f(e^{-2 \pi i t})\,dt = \int_0^1 f(x) f(x^{-1})\big|_{x = e^{2 \pi i t}}\,dt .$$ Now assume that $f(x) = g(x) h(x)$. Then, since $f(x) f(x^{-1}) = \bigl(g(x)h(x^{-1})\bigr)\bigl(g(x^{-1})h(x)\bigr)$, $G(x) := g(x) h(x^{-1})$ satisfies $\|G\| = \|f\|$ and $G(x) G(x^{-1}) = f(x) f(x^{-1})$.</p> <p>Now we consider $f(x) = x^n - 7 x^2 + 1$; then $\|f\| = 51$. If $f = g h$ as above, then write $G(x) = \pm x^m(1 + a_1 x + a_2 x^2 + \ldots)$ and $G(x^{-1}) = \pm x^l(1 + b_1 x + b_2 x^2 + \ldots)$. The relation $G(x) G(x^{-1}) = f(x) f(x^{-1})$ translates into (equality of signs and) $$(1 + a_1 x + \ldots)(1 + b_1 x + \ldots) = 1 - 7 x^2 + O(x^{n-2}).$$ Assuming that $n &gt; 40$ and considering terms up to $x^{20}$, one can check (see below) that the only solution such that $a_1^2 + a_2^2 + \ldots + a_{20}^2 + b_1^2 + b_2^2 + \ldots + b_{20}^2\le 49$ is, up to the substitution $x \leftarrow x^{-1}$, given by $1 + a_1 x + \ldots = 1 - 7x^2 + O(x^{21})$, $1 + b_1 x + \ldots = 1 + O(x^{21})$. Since the $-7$ (together with the leading and trailing 1) exhausts our allowance for the sum of squares of the coefficients, all other coefficients must be zero, and we obtain that $G(x) = \pm x^a f(x)$ or $G(x) = \pm x^a f(x^{-1})$. Modulo interchanging $g$ and $h$, we can assume that $g(x) h(x^{-1}) = \pm x^a f(x)$, so $h(x^{-1}) = \pm x^a f(x)/g(x) = \pm x^a h(x)$, and $x^{\deg h} h(x^{-1})$ divides $f(x)$. This implies that $h(x)$ divides $x^n f(x^{-1})$, so $h(x)$ must divide $$f(x) - x^n f(x^{-1}) = 7 x^2 (x^{n-4} - 1).$$ So $h$ also divides $$f(x) - x^4 (x^{n-4} - 1) = x^4 - 7 x^2 + 1 = (x^2-3x+1)(x^2+3x+1).$$ Since $h$ also divides $f$, it must divide the difference $x^n - x^4$, which for $n \neq 4$ it clearly doesn't, since the quartic has no roots of absolute value 0 or 1; contradiction.</p> <p>The argument shows that $x^n - 7 x^2 + 1$ is irreducible for $n &gt; 40$; for smaller $n$, we can ask the Computer Algebra System we trust. This gives:</p> <p><strong>Theorem.</strong> $x^n - 7 x^2 + 1$ is irreducible over $\mathbb Q$ for all positive integers $n$ except $n=4$.</p> <p><strong>ADDED 2017-01-08:</strong> After re-checking the computations, I realized that there was a small mistake that ruled out some partial solutions prematurely. It looks like one needs to consider terms up to $x^{20}$. <a href="http://www.mathe2.uni-bayreuth.de/stoll/MO/MO-258914.magma">Here</a> is a file with MAGMA code that verifies that there are no other solutions.</p> Answer 5: <p>Not an answer, but an observation, which, if proved, will probably give an answer.</p> <p>For all $k$ that I have computed, your polynomial has two roots smaller than $1$ in absolute value (for large $k$ these converge to $\pm\frac{1}{\sqrt{7}}.$), and the rest of the roots are larger than $1$ in absolute value. Let's assume this is true. Since the constant term of a factor has to have absolute value $1$ each factor should have one of these roots, so there are at most $2$ irreducible factors.</p> Answer 6: <p>EDIT: I'm afraid that this answer is wrong and I hope you have some way of "unaccepting" it. The reason is that when taking conjugates of $\frac{1}{\sqrt{7}-\beta_1}$ we're allowed to freely change the sign of $\sqrt{7}$ (unless $\sqrt{7}$ is contained in the field generated by $\beta_1$, which seems rather unlikely). And $\frac{1}{-\sqrt{7}-\beta_1}$ is greater than $1$ in absolute value. I found this mistake only when I noticed that my "proof" gives that $x^4-7x^2+1=(x^2+3x+1)(x^2-3x+1)$ is irreducible. It should work with an integer square in place of $7$. Maybe someone can make it work also for $7$? I apologize for raising false hopes!</p> <p>The question is equivalent (by replacing $x$ with $\frac{1}{x}$) to asking about the irreducibility of $A(x) = x^{2k+1}-7x^{2k-1}+1 = x^{2k-1}(x^2-7)+1$ over $\mathbb{Q}$. An application of Rouché's theorem shows that $A(x)$ has exactly $2k-1$ zeroes (counted with multiplicity) of absolute value less than $1$. There remain two real zeroes $\beta_1 &lt; -1$ and $\beta_2 &gt; 1$. One can deduce from the equation they satisfy that $|\beta_i - (-1)^i\sqrt{7}| \leq \frac{1}{\sqrt{7}+1}$ ($i=1,2$).</p> <p>Now suppose that $A$ factorizes as $A = BC$ with monic integer polynomials $B, C$. Wlog $B(\beta_1) = 0$. If $B(\beta_2) \neq 0$, then $$\beta_1^{2k-1}(\beta_1+\sqrt{7}) = \frac{1}{\sqrt{7}-\beta_1}$$ is an algebraic integer, all conjugates of which are less than $1$ in absolute value (the right hand side is less than $\frac{1}{2\sqrt{7}-\frac{1}{\sqrt{7}+1}}&lt;1$ and all its conjugates are of absolute value at most $\frac{1}{\sqrt{7}-1}&lt;1$).</p> <p>So it would have to be $0$, which is absurd. So $B(\beta_2)=0$. But now all roots of $C$ are less than $1$ in absolute value, while its constant term is $1$. It follows that $C=1$ is constant and so $A$ is irreducible.</p> Answer 7: <p>Here is a proof, based on a trick that can be used to prove that $x^n + x + 1$ is irreducible when $n \not\equiv 2 \bmod 3$.</p> <p>We work with Laurent polynomials in $R = \mathbb Z[x,x^{-1}]$; note that $R$ has unit group $R^\times = \pm x^{\mathbb Z}$. We observe that for $f \in R$, the sum of the squares of the coefficients is given by $$\|f\| = \int_0^1 |f(e^{2 \pi i t})|^2\,dt = \int_0^1 f(e^{2 \pi i t}) f(e^{-2 \pi i t})\,dt = \int_0^1 f(x) f(x^{-1})\big|_{x = e^{2 \pi i t}}\,dt .$$ Now assume that $f(x) = g(x) h(x)$. Then, since $f(x) f(x^{-1}) = \bigl(g(x)h(x^{-1})\bigr)\bigl(g(x^{-1})h(x)\bigr)$, $G(x) := g(x) h(x^{-1})$ satisfies $\|G\| = \|f\|$ and $G(x) G(x^{-1}) = f(x) f(x^{-1})$.</p> <p>Now we consider $f(x) = x^n - 7 x^2 + 1$; then $\|f\| = 51$. If $f = g h$ as above, then write $G(x) = \pm x^m(1 + a_1 x + a_2 x^2 + \ldots)$ and $G(x^{-1}) = \pm x^l(1 + b_1 x + b_2 x^2 + \ldots)$. The relation $G(x) G(x^{-1}) = f(x) f(x^{-1})$ translates into (equality of signs and) $$(1 + a_1 x + \ldots)(1 + b_1 x + \ldots) = 1 - 7 x^2 + O(x^{n-2}).$$ Assuming that $n &gt; 40$ and considering terms up to $x^{20}$, one can check (see below) that the only solution such that $a_1^2 + a_2^2 + \ldots + a_{20}^2 + b_1^2 + b_2^2 + \ldots + b_{20}^2\le 49$ is, up to the substitution $x \leftarrow x^{-1}$, given by $1 + a_1 x + \ldots = 1 - 7x^2 + O(x^{21})$, $1 + b_1 x + \ldots = 1 + O(x^{21})$. Since the $-7$ (together with the leading and trailing 1) exhausts our allowance for the sum of squares of the coefficients, all other coefficients must be zero, and we obtain that $G(x) = \pm x^a f(x)$ or $G(x) = \pm x^a f(x^{-1})$. Modulo interchanging $g$ and $h$, we can assume that $g(x) h(x^{-1}) = \pm x^a f(x)$, so $h(x^{-1}) = \pm x^a f(x)/g(x) = \pm x^a h(x)$, and $x^{\deg h} h(x^{-1})$ divides $f(x)$. This implies that $h(x)$ divides $x^n f(x^{-1})$, so $h(x)$ must divide $$f(x) - x^n f(x^{-1}) = 7 x^2 (x^{n-4} - 1).$$ So $h$ also divides $$f(x) - x^4 (x^{n-4} - 1) = x^4 - 7 x^2 + 1 = (x^2-3x+1)(x^2+3x+1).$$ Since $h$ also divides $f$, it must divide the difference $x^n - x^4$, which for $n \neq 4$ it clearly doesn't, since the quartic has no roots of absolute value 0 or 1; contradiction.</p> <p>The argument shows that $x^n - 7 x^2 + 1$ is irreducible for $n &gt; 40$; for smaller $n$, we can ask the Computer Algebra System we trust. This gives:</p> <p><strong>Theorem.</strong> $x^n - 7 x^2 + 1$ is irreducible over $\mathbb Q$ for all positive integers $n$ except $n=4$.</p> <p><strong>ADDED 2017-01-08:</strong> After re-checking the computations, I realized that there was a small mistake that ruled out some partial solutions prematurely. It looks like one needs to consider terms up to $x^{20}$. <a href="http://www.mathe2.uni-bayreuth.de/stoll/MO/MO-258914.magma">Here</a> is a file with MAGMA code that verifies that there are no other solutions.</p> Answer 8: <p>Not an answer, but an observation, which, if proved, will probably give an answer.</p> <p>For all $k$ that I have computed, your polynomial has two roots smaller than $1$ in absolute value (for large $k$ these converge to $\pm\frac{1}{\sqrt{7}}.$), and the rest of the roots are larger than $1$ in absolute value. Let's assume this is true. Since the constant term of a factor has to have absolute value $1$ each factor should have one of these roots, so there are at most $2$ irreducible factors.</p> Answer 9: <p>EDIT: I'm afraid that this answer is wrong and I hope you have some way of "unaccepting" it. The reason is that when taking conjugates of $\frac{1}{\sqrt{7}-\beta_1}$ we're allowed to freely change the sign of $\sqrt{7}$ (unless $\sqrt{7}$ is contained in the field generated by $\beta_1$, which seems rather unlikely). And $\frac{1}{-\sqrt{7}-\beta_1}$ is greater than $1$ in absolute value. I found this mistake only when I noticed that my "proof" gives that $x^4-7x^2+1=(x^2+3x+1)(x^2-3x+1)$ is irreducible. It should work with an integer square in place of $7$. Maybe someone can make it work also for $7$? I apologize for raising false hopes!</p> <p>The question is equivalent (by replacing $x$ with $\frac{1}{x}$) to asking about the irreducibility of $A(x) = x^{2k+1}-7x^{2k-1}+1 = x^{2k-1}(x^2-7)+1$ over $\mathbb{Q}$. An application of Rouché's theorem shows that $A(x)$ has exactly $2k-1$ zeroes (counted with multiplicity) of absolute value less than $1$. There remain two real zeroes $\beta_1 &lt; -1$ and $\beta_2 &gt; 1$. One can deduce from the equation they satisfy that $|\beta_i - (-1)^i\sqrt{7}| \leq \frac{1}{\sqrt{7}+1}$ ($i=1,2$).</p> <p>Now suppose that $A$ factorizes as $A = BC$ with monic integer polynomials $B, C$. Wlog $B(\beta_1) = 0$. If $B(\beta_2) \neq 0$, then $$\beta_1^{2k-1}(\beta_1+\sqrt{7}) = \frac{1}{\sqrt{7}-\beta_1}$$ is an algebraic integer, all conjugates of which are less than $1$ in absolute value (the right hand side is less than $\frac{1}{2\sqrt{7}-\frac{1}{\sqrt{7}+1}}&lt;1$ and all its conjugates are of absolute value at most $\frac{1}{\sqrt{7}-1}&lt;1$).</p> <p>So it would have to be $0$, which is absurd. So $B(\beta_2)=0$. But now all roots of $C$ are less than $1$ in absolute value, while its constant term is $1$. It follows that $C=1$ is constant and so $A$ is irreducible.</p> Answer 10: <p>Here is a proof, based on a trick that can be used to prove that $x^n + x + 1$ is irreducible when $n \not\equiv 2 \bmod 3$.</p> <p>We work with Laurent polynomials in $R = \mathbb Z[x,x^{-1}]$; note that $R$ has unit group $R^\times = \pm x^{\mathbb Z}$. We observe that for $f \in R$, the sum of the squares of the coefficients is given by $$\|f\| = \int_0^1 |f(e^{2 \pi i t})|^2\,dt = \int_0^1 f(e^{2 \pi i t}) f(e^{-2 \pi i t})\,dt = \int_0^1 f(x) f(x^{-1})\big|_{x = e^{2 \pi i t}}\,dt .$$ Now assume that $f(x) = g(x) h(x)$. Then, since $f(x) f(x^{-1}) = \bigl(g(x)h(x^{-1})\bigr)\bigl(g(x^{-1})h(x)\bigr)$, $G(x) := g(x) h(x^{-1})$ satisfies $\|G\| = \|f\|$ and $G(x) G(x^{-1}) = f(x) f(x^{-1})$.</p> <p>Now we consider $f(x) = x^n - 7 x^2 + 1$; then $\|f\| = 51$. If $f = g h$ as above, then write $G(x) = \pm x^m(1 + a_1 x + a_2 x^2 + \ldots)$ and $G(x^{-1}) = \pm x^l(1 + b_1 x + b_2 x^2 + \ldots)$. The relation $G(x) G(x^{-1}) = f(x) f(x^{-1})$ translates into (equality of signs and) $$(1 + a_1 x + \ldots)(1 + b_1 x + \ldots) = 1 - 7 x^2 + O(x^{n-2}).$$ Assuming that $n &gt; 40$ and considering terms up to $x^{20}$, one can check (see below) that the only solution such that $a_1^2 + a_2^2 + \ldots + a_{20}^2 + b_1^2 + b_2^2 + \ldots + b_{20}^2\le 49$ is, up to the substitution $x \leftarrow x^{-1}$, given by $1 + a_1 x + \ldots = 1 - 7x^2 + O(x^{21})$, $1 + b_1 x + \ldots = 1 + O(x^{21})$. Since the $-7$ (together with the leading and trailing 1) exhausts our allowance for the sum of squares of the coefficients, all other coefficients must be zero, and we obtain that $G(x) = \pm x^a f(x)$ or $G(x) = \pm x^a f(x^{-1})$. Modulo interchanging $g$ and $h$, we can assume that $g(x) h(x^{-1}) = \pm x^a f(x)$, so $h(x^{-1}) = \pm x^a f(x)/g(x) = \pm x^a h(x)$, and $x^{\deg h} h(x^{-1})$ divides $f(x)$. This implies that $h(x)$ divides $x^n f(x^{-1})$, so $h(x)$ must divide $$f(x) - x^n f(x^{-1}) = 7 x^2 (x^{n-4} - 1).$$ So $h$ also divides $$f(x) - x^4 (x^{n-4} - 1) = x^4 - 7 x^2 + 1 = (x^2-3x+1)(x^2+3x+1).$$ Since $h$ also divides $f$, it must divide the difference $x^n - x^4$, which for $n \neq 4$ it clearly doesn't, since the quartic has no roots of absolute value 0 or 1; contradiction.</p> <p>The argument shows that $x^n - 7 x^2 + 1$ is irreducible for $n &gt; 40$; for smaller $n$, we can ask the Computer Algebra System we trust. This gives:</p> <p><strong>Theorem.</strong> $x^n - 7 x^2 + 1$ is irreducible over $\mathbb Q$ for all positive integers $n$ except $n=4$.</p> <p><strong>ADDED 2017-01-08:</strong> After re-checking the computations, I realized that there was a small mistake that ruled out some partial solutions prematurely. It looks like one needs to consider terms up to $x^{20}$. <a href="http://www.mathe2.uni-bayreuth.de/stoll/MO/MO-258914.magma">Here</a> is a file with MAGMA code that verifies that there are no other solutions.</p> Answer 11: <p>Not an answer, but an observation, which, if proved, will probably give an answer.</p> <p>For all $k$ that I have computed, your polynomial has two roots smaller than $1$ in absolute value (for large $k$ these converge to $\pm\frac{1}{\sqrt{7}}.$), and the rest of the roots are larger than $1$ in absolute value. Let's assume this is true. Since the constant term of a factor has to have absolute value $1$ each factor should have one of these roots, so there are at most $2$ irreducible factors.</p> Answer 12: <p>EDIT: I'm afraid that this answer is wrong and I hope you have some way of "unaccepting" it. The reason is that when taking conjugates of $\frac{1}{\sqrt{7}-\beta_1}$ we're allowed to freely change the sign of $\sqrt{7}$ (unless $\sqrt{7}$ is contained in the field generated by $\beta_1$, which seems rather unlikely). And $\frac{1}{-\sqrt{7}-\beta_1}$ is greater than $1$ in absolute value. I found this mistake only when I noticed that my "proof" gives that $x^4-7x^2+1=(x^2+3x+1)(x^2-3x+1)$ is irreducible. It should work with an integer square in place of $7$. Maybe someone can make it work also for $7$? I apologize for raising false hopes!</p> <p>The question is equivalent (by replacing $x$ with $\frac{1}{x}$) to asking about the irreducibility of $A(x) = x^{2k+1}-7x^{2k-1}+1 = x^{2k-1}(x^2-7)+1$ over $\mathbb{Q}$. An application of Rouché's theorem shows that $A(x)$ has exactly $2k-1$ zeroes (counted with multiplicity) of absolute value less than $1$. There remain two real zeroes $\beta_1 &lt; -1$ and $\beta_2 &gt; 1$. One can deduce from the equation they satisfy that $|\beta_i - (-1)^i\sqrt{7}| \leq \frac{1}{\sqrt{7}+1}$ ($i=1,2$).</p> <p>Now suppose that $A$ factorizes as $A = BC$ with monic integer polynomials $B, C$. Wlog $B(\beta_1) = 0$. If $B(\beta_2) \neq 0$, then $$\beta_1^{2k-1}(\beta_1+\sqrt{7}) = \frac{1}{\sqrt{7}-\beta_1}$$ is an algebraic integer, all conjugates of which are less than $1$ in absolute value (the right hand side is less than $\frac{1}{2\sqrt{7}-\frac{1}{\sqrt{7}+1}}&lt;1$ and all its conjugates are of absolute value at most $\frac{1}{\sqrt{7}-1}&lt;1$).</p> <p>So it would have to be $0$, which is absurd. So $B(\beta_2)=0$. But now all roots of $C$ are less than $1$ in absolute value, while its constant term is $1$. It follows that $C=1$ is constant and so $A$ is irreducible.</p>
Nice applications for Schwartz distributions
259,834
26
https://mathoverflow.net/questions/259834/nice-applications-for-schwartz-distributions
<p>I am to teach a second year grad course in analysis with focus on Schwartz distributions. Among the core topics I intend to cover are:</p> <ol> <li><p>Some multilinear algebra including the Kernel Theorem and Volterra composition,</p></li> <li><p>Some Fourier analysis including the Bochner-Schwartz Theorem,</p></li> <li><p>An introduction to wavelets with a view to structure theorems for spaces of distributions or function spaces,</p></li> <li><p>Probability theory on spaces of distributions including the L&eacute;vy Continuity Theorem,</p></li> <li><p>A study of homogeneous distributions and elementary solutions to linear PDEs.</p></li> </ol> <p><strong>My question is:</strong> What "cool topics/applications" would it be nice to include in such a course? I am particularly interested in examples with a high return on investment, i.e., that would not take too long to cover yet would provide the students with valuable tools for eventually a future research career in analysis. Please provide references where I can learn more about your suggestions. I would like some variety if possible. I got suggestions pertaining to probability, PDEs and mathematical physics, but it would be nice to get apps related to other areas of math.</p>
[ { "answer_id": 260634, "score": 27, "text": "<p>Unsurprisingly, the topics that occur to me have various connections to number theory (and related harmonic analysis) (and unclear to me what might have already been done in your course...):</p>\n\n<p>EDIT: inserted some links... EDIT-EDIT: one more...</p>\n\n<ul>\n<li><p>Genuinely distributional proof of Poisson summation: <a href=\"http://www.math.umn.edu/~garrett/m/fun/poisson.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/poisson.pdf</a></p></li>\n<li><p>Meromorphic continuation of distributions $|x|^s$ and ${\\mathrm sgn}(x)\\cdot |x|^s$ <a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf</a></p></li>\n<li><p>(Fancier version of the previous: mero cont'n of $|\\det x|^s$ on $n\\times n$ matrices (if this is interesting, I have some notes, or maybe it's a fun exercise). In particular, stimulated by a math-overflow question of A. Braverman some time ago, there <em>are</em> equivariant distributions (e.g., on two-by-two matrices) such that both they and their Fourier transform are supported on singular matrices... Wacky! <a href=\"http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf</a></p></li>\n<li><p>Decomposition of $L^2(A)$ for compact abelian topological groups $A$ (by Hilbert-Schmidt, hence compact, operators). <a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf</a></p></li>\n<li><p>Reconsideration of Sturm-Liouville problems (with reasonable hypotheses), to really <em>prove</em> things that are ... ahem... \"suggested\" in usual more-naive discussions.</p></li>\n<li><p>Quadratic reciprocity over number fields (and function fields not of char=2) ... cf. <a href=\"http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf</a>\n(This presumes Poisson summation for $\\mathbb A/k$...)</p></li>\n<li><p>Explanation that Schwartz' kernel theorem is a corollary of a Cartan-Eilenberg adjunction (between $\\otimes$ and $\\mathrm{Hom(,-)}$), when we know that there are genuine (i.e., categorically correct) tensor products for \"nuclear Frechet spaces\", ... which leads to the issue of suitable notions of the latter. <a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf</a></p></li>\n<li><p>The idea that termwise differentiation of Fourier series is \"<em>always</em> ok\" (with coefs that grow at most polynomially), if/when the outcome is interpreted as lying in a suitable Sobolev space on the circle. And that polynomial-growth-coefficiented Fourier series _always_converge_... if only in a suitable Sobolev space.\n<a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf</a></p></li>\n</ul>\n\n<p>** Possibility of ranting about the limitations of pointwise convergence... especially when placed in contrast to convergence in Sobolev spaces...</p>\n\n<ul>\n<li><p>Use of Snake Lemma to talk about mero cont'n of the Gamma function, via mero cont'n of $|x|^s$. :) <a href=\"http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf</a></p></li>\n<li><p>Peetre's theorem that any (not necessarily continuous!) linear functional on test functions that does not increase support is a differential operator. (I have a note on this, which may be more palatable to beginners than Peetre's paper.) <a href=\"http://www.math.umn.edu/~garrett/m/v/peetre.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/peetre.pdf</a></p></li>\n<li><p>Uniqueness of invariant functionals... As the easiest case (which is easy, but cognitive-dissonance-provoking, in my experience), proving that $u'=0$ for a distribution $u$ implies that $u$ is (integration-against) a constant. (Maybe you'd do this along the way...) <a href=\"http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf</a></p></li>\n<li><p>... this is not a stand-alone topic, but: the usual discussions of pseudo-differential operators (e.g., \"symbols\") somehow shrink from talking about <em>quotients</em> of TVS's in a grown-up way... If that hadn't been done earlier, and/or people had a (reasonable!) feeling of discomfort about the usual style of chatter in the psi-DO world, perhaps this could be happy-making.</p></li>\n<li><p>Meromorphic/holomorphic vector-valued functions (cf. Grothendieck c. 1953-4, and also various of my notes) e.g., meromorphic families of distributions... E.g., the $|x|^s$ family on $\\mathbb R$ has <em>residues</em> which are the even-order derivatives of $\\delta$, and ${\\mathrm {sgn}}(x)\\cdot |x|^s$ has as residues the odd-order derivatives of $\\delta$. <a href=\"http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf</a>\nDepending on context, there are somewhat-fancier things that I do find entertaining and also useful. Comments/correspondence are welcome.</p></li>\n</ul>\n" }, { "answer_id": 259874, "score": 12, "text": "<p>Two (<strong>edit:</strong> now four) not-so-usual examples come to my mind:</p>\n\n<ul>\n<li>There is the proof of the <em>central limit theorem</em> using Fourier analysis, as done in Chapter 7 of Hörmander's book. It's a cornerstone of probability theory and the Fourier-analytic proof presented by Hörmander is a nice showcase of the power of distribution theory: one gets a conceptually deep and useful theorem in full generality with relatively little effort;</li>\n<li>Something I have been meaning to try is to work out some \"mathematical toy examples\" of continuum limits of distributions supported on a lattice, as done e.g. in Chapter 15 of the book by R. Fernández, J. Fröhlich and A.D. Sokal, <em>Random Walks, Critical Phenomena, and Triviality in Quantum Field Theory</em> (Springer, 1992). I do not know if it is a feasible task, but since you are a specialist is constructive QFT, you should be able to do it... <strong>Edit 3:</strong> a classical and rather charming example of this has just came to my attention: the so-called <em>Bernstein polynomials</em>, used by their namesake Sergei N. Bernstein (the same mathematician who found the family of inequalities - also named after him - for smooth functions with compactly supported Fourier transform which underlie the Littlewood-Paley theory discussed in my Edit 1 below) to provide a probabilistic proof of the Weierstrass approximation theorem through the weak law of large numbers. Taking formal adjoints of the sequence of Bernstein polynomial operators provides a general lattice approximation for distributions on compact rectangular domains. A rather extensive discussion can be found in the short and beautiful book of the same title by G. G. Lorentz (the mathematician who invented the so-called Lorentz interpolation spaces, not the Dutch physicist).</li>\n<li><strong>(Edit 1)</strong> A third one, which I am particularly fond of, is the dyadic Fourier analysis of Littlewood and Paley, which is strongly connected to point (3). The connection to renormalization group ideas is quite obvious, and it has become a standard tool for obtaining important inequalities in nonlinear analysis (Gagliardo-Nirenberg, Moser, Schauder), all the way to the monumental Nash-Moser implicit function theorem (as shown by Hörmander - him again - in the 70's). A nice pedagogical exposition of these ideas can be found the book by S. Alinhac and P. Gérard, <em>Pseudo-Differential Operators and the Nash-Moser Theorem</em> (AMS, 2007) - just watch out for the numerous typographic errors!</li>\n<li><strong>(Edit 2)</strong> A fourth one, which is not actually unusual in itself but rather in its standpoint, is that many interesting examples of distributions (principal values, <em>partes finies</em>, extensions of homogeneous distributions to the origin, etc.) are obtained as extensions of distributions initially defined in (often finite-codimensional) subspaces of test functions, collectively called <em>renormalizations</em>. What would be interesting, also regarding applications to physics, is to treat such examples from this unified viewpoint. Indeed, as it can be seen e.g. by (Bogoliubov-Parasiuk-)Hepp's and Epstein-Glaser's treatments of perturbative renormalization in QFT, the latter really amounts to renormalizing distributions in the above sense. As a famous dictum of K. Hepp goes (<em>Proof of the Bogoliubov-Parasiuk Theorem on Renormalization</em>, Commun. Math. Phys. <strong>2</strong> (1966) 301-326):\n\n<blockquote>\n <p>\"The renormalization theory (of Dyson) is in this framework a constructive \n form of the Hahn-Banach theorem.\" (pp. 302 - parentheses mine)</p>\n</blockquote></li>\n</ul>\n" }, { "answer_id": 259864, "score": 6, "text": "<p>One of my favourite applications of 'basic' distribution theory, which actually requires most of the tools in your list to be fully apreciated, is the Malgrange--Ehrenpreis Theorem on the local solvability of arbitrary constant coefficients PDE. There are a few proofs, but the proof in M.Taylor's book on Pseudodifferential Operators (last Section of first Chapter) is especially suited for your course I think. It is a substantial theorem though, so it might or might be not a good idea to include it, depending on the length and size of the course.</p>\n" }, { "answer_id": 289039, "score": 3, "text": "<p>No so much as applications, but I would love to learn about the basics of hyperfunctions (of the Japanese school) in such a course, if only in one dimension. Specifically I think it gives an easier way to define the wavefront set, which is what's required to explain why we can't square the delta function (referring to a comment above). See also the question <a href=\"https://mathoverflow.net/questions/48067/is-square-of-delta-function-defined-somewhere\">Is square of Delta function defined somewhere?</a></p>\n\n<p>For what it's worth, I found the book</p>\n\n<ul>\n<li>Kato &amp; Struppa, <em>Fundamentals of Algebraic Microlocal Analysis</em></li>\n</ul>\n\n<p>particularly well written. The article in EoM is also great</p>\n\n<p><a href=\"https://www.encyclopediaofmath.org/index.php/Hyperfunction\" rel=\"nofollow noreferrer\">https://www.encyclopediaofmath.org/index.php/Hyperfunction</a></p>\n" }, { "answer_id": 298578, "score": 1, "text": "<p>Let me mention two nice applications to Sobolev spaces:</p>\n\n<p>The first result shows that if $f\\in \\mathcal{D}'$ and $\\nabla f\\in L^p_{\\rm loc}$, then $f\\in L^p_{\\rm loc}$. For a short proof of this classical result see\n<a href=\"https://mathoverflow.net/a/296464/121665\">https://mathoverflow.net/a/296464/121665</a></p>\n\n<p>For applications of this result to higher order Sobolev spaces see:<a href=\"https://mathoverflow.net/a/297392/121665\">https://mathoverflow.net/a/297392/121665</a></p>\n" }, { "answer_id": 260634, "score": 27, "text": "<p>Unsurprisingly, the topics that occur to me have various connections to number theory (and related harmonic analysis) (and unclear to me what might have already been done in your course...):</p>\n\n<p>EDIT: inserted some links... EDIT-EDIT: one more...</p>\n\n<ul>\n<li><p>Genuinely distributional proof of Poisson summation: <a href=\"http://www.math.umn.edu/~garrett/m/fun/poisson.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/poisson.pdf</a></p></li>\n<li><p>Meromorphic continuation of distributions $|x|^s$ and ${\\mathrm sgn}(x)\\cdot |x|^s$ <a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf</a></p></li>\n<li><p>(Fancier version of the previous: mero cont'n of $|\\det x|^s$ on $n\\times n$ matrices (if this is interesting, I have some notes, or maybe it's a fun exercise). In particular, stimulated by a math-overflow question of A. Braverman some time ago, there <em>are</em> equivariant distributions (e.g., on two-by-two matrices) such that both they and their Fourier transform are supported on singular matrices... Wacky! <a href=\"http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf</a></p></li>\n<li><p>Decomposition of $L^2(A)$ for compact abelian topological groups $A$ (by Hilbert-Schmidt, hence compact, operators). <a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf</a></p></li>\n<li><p>Reconsideration of Sturm-Liouville problems (with reasonable hypotheses), to really <em>prove</em> things that are ... ahem... \"suggested\" in usual more-naive discussions.</p></li>\n<li><p>Quadratic reciprocity over number fields (and function fields not of char=2) ... cf. <a href=\"http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf</a>\n(This presumes Poisson summation for $\\mathbb A/k$...)</p></li>\n<li><p>Explanation that Schwartz' kernel theorem is a corollary of a Cartan-Eilenberg adjunction (between $\\otimes$ and $\\mathrm{Hom(,-)}$), when we know that there are genuine (i.e., categorically correct) tensor products for \"nuclear Frechet spaces\", ... which leads to the issue of suitable notions of the latter. <a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf</a></p></li>\n<li><p>The idea that termwise differentiation of Fourier series is \"<em>always</em> ok\" (with coefs that grow at most polynomially), if/when the outcome is interpreted as lying in a suitable Sobolev space on the circle. And that polynomial-growth-coefficiented Fourier series _always_converge_... if only in a suitable Sobolev space.\n<a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf</a></p></li>\n</ul>\n\n<p>** Possibility of ranting about the limitations of pointwise convergence... especially when placed in contrast to convergence in Sobolev spaces...</p>\n\n<ul>\n<li><p>Use of Snake Lemma to talk about mero cont'n of the Gamma function, via mero cont'n of $|x|^s$. :) <a href=\"http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf</a></p></li>\n<li><p>Peetre's theorem that any (not necessarily continuous!) linear functional on test functions that does not increase support is a differential operator. (I have a note on this, which may be more palatable to beginners than Peetre's paper.) <a href=\"http://www.math.umn.edu/~garrett/m/v/peetre.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/peetre.pdf</a></p></li>\n<li><p>Uniqueness of invariant functionals... As the easiest case (which is easy, but cognitive-dissonance-provoking, in my experience), proving that $u'=0$ for a distribution $u$ implies that $u$ is (integration-against) a constant. (Maybe you'd do this along the way...) <a href=\"http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf</a></p></li>\n<li><p>... this is not a stand-alone topic, but: the usual discussions of pseudo-differential operators (e.g., \"symbols\") somehow shrink from talking about <em>quotients</em> of TVS's in a grown-up way... If that hadn't been done earlier, and/or people had a (reasonable!) feeling of discomfort about the usual style of chatter in the psi-DO world, perhaps this could be happy-making.</p></li>\n<li><p>Meromorphic/holomorphic vector-valued functions (cf. Grothendieck c. 1953-4, and also various of my notes) e.g., meromorphic families of distributions... E.g., the $|x|^s$ family on $\\mathbb R$ has <em>residues</em> which are the even-order derivatives of $\\delta$, and ${\\mathrm {sgn}}(x)\\cdot |x|^s$ has as residues the odd-order derivatives of $\\delta$. <a href=\"http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf</a>\nDepending on context, there are somewhat-fancier things that I do find entertaining and also useful. Comments/correspondence are welcome.</p></li>\n</ul>\n" }, { "answer_id": 259874, "score": 12, "text": "<p>Two (<strong>edit:</strong> now four) not-so-usual examples come to my mind:</p>\n\n<ul>\n<li>There is the proof of the <em>central limit theorem</em> using Fourier analysis, as done in Chapter 7 of Hörmander's book. It's a cornerstone of probability theory and the Fourier-analytic proof presented by Hörmander is a nice showcase of the power of distribution theory: one gets a conceptually deep and useful theorem in full generality with relatively little effort;</li>\n<li>Something I have been meaning to try is to work out some \"mathematical toy examples\" of continuum limits of distributions supported on a lattice, as done e.g. in Chapter 15 of the book by R. Fernández, J. Fröhlich and A.D. Sokal, <em>Random Walks, Critical Phenomena, and Triviality in Quantum Field Theory</em> (Springer, 1992). I do not know if it is a feasible task, but since you are a specialist is constructive QFT, you should be able to do it... <strong>Edit 3:</strong> a classical and rather charming example of this has just came to my attention: the so-called <em>Bernstein polynomials</em>, used by their namesake Sergei N. Bernstein (the same mathematician who found the family of inequalities - also named after him - for smooth functions with compactly supported Fourier transform which underlie the Littlewood-Paley theory discussed in my Edit 1 below) to provide a probabilistic proof of the Weierstrass approximation theorem through the weak law of large numbers. Taking formal adjoints of the sequence of Bernstein polynomial operators provides a general lattice approximation for distributions on compact rectangular domains. A rather extensive discussion can be found in the short and beautiful book of the same title by G. G. Lorentz (the mathematician who invented the so-called Lorentz interpolation spaces, not the Dutch physicist).</li>\n<li><strong>(Edit 1)</strong> A third one, which I am particularly fond of, is the dyadic Fourier analysis of Littlewood and Paley, which is strongly connected to point (3). The connection to renormalization group ideas is quite obvious, and it has become a standard tool for obtaining important inequalities in nonlinear analysis (Gagliardo-Nirenberg, Moser, Schauder), all the way to the monumental Nash-Moser implicit function theorem (as shown by Hörmander - him again - in the 70's). A nice pedagogical exposition of these ideas can be found the book by S. Alinhac and P. Gérard, <em>Pseudo-Differential Operators and the Nash-Moser Theorem</em> (AMS, 2007) - just watch out for the numerous typographic errors!</li>\n<li><strong>(Edit 2)</strong> A fourth one, which is not actually unusual in itself but rather in its standpoint, is that many interesting examples of distributions (principal values, <em>partes finies</em>, extensions of homogeneous distributions to the origin, etc.) are obtained as extensions of distributions initially defined in (often finite-codimensional) subspaces of test functions, collectively called <em>renormalizations</em>. What would be interesting, also regarding applications to physics, is to treat such examples from this unified viewpoint. Indeed, as it can be seen e.g. by (Bogoliubov-Parasiuk-)Hepp's and Epstein-Glaser's treatments of perturbative renormalization in QFT, the latter really amounts to renormalizing distributions in the above sense. As a famous dictum of K. Hepp goes (<em>Proof of the Bogoliubov-Parasiuk Theorem on Renormalization</em>, Commun. Math. Phys. <strong>2</strong> (1966) 301-326):\n\n<blockquote>\n <p>\"The renormalization theory (of Dyson) is in this framework a constructive \n form of the Hahn-Banach theorem.\" (pp. 302 - parentheses mine)</p>\n</blockquote></li>\n</ul>\n" }, { "answer_id": 259864, "score": 6, "text": "<p>One of my favourite applications of 'basic' distribution theory, which actually requires most of the tools in your list to be fully apreciated, is the Malgrange--Ehrenpreis Theorem on the local solvability of arbitrary constant coefficients PDE. There are a few proofs, but the proof in M.Taylor's book on Pseudodifferential Operators (last Section of first Chapter) is especially suited for your course I think. It is a substantial theorem though, so it might or might be not a good idea to include it, depending on the length and size of the course.</p>\n" }, { "answer_id": 289039, "score": 3, "text": "<p>No so much as applications, but I would love to learn about the basics of hyperfunctions (of the Japanese school) in such a course, if only in one dimension. Specifically I think it gives an easier way to define the wavefront set, which is what's required to explain why we can't square the delta function (referring to a comment above). See also the question <a href=\"https://mathoverflow.net/questions/48067/is-square-of-delta-function-defined-somewhere\">Is square of Delta function defined somewhere?</a></p>\n\n<p>For what it's worth, I found the book</p>\n\n<ul>\n<li>Kato &amp; Struppa, <em>Fundamentals of Algebraic Microlocal Analysis</em></li>\n</ul>\n\n<p>particularly well written. The article in EoM is also great</p>\n\n<p><a href=\"https://www.encyclopediaofmath.org/index.php/Hyperfunction\" rel=\"nofollow noreferrer\">https://www.encyclopediaofmath.org/index.php/Hyperfunction</a></p>\n" }, { "answer_id": 298578, "score": 1, "text": "<p>Let me mention two nice applications to Sobolev spaces:</p>\n\n<p>The first result shows that if $f\\in \\mathcal{D}'$ and $\\nabla f\\in L^p_{\\rm loc}$, then $f\\in L^p_{\\rm loc}$. For a short proof of this classical result see\n<a href=\"https://mathoverflow.net/a/296464/121665\">https://mathoverflow.net/a/296464/121665</a></p>\n\n<p>For applications of this result to higher order Sobolev spaces see:<a href=\"https://mathoverflow.net/a/297392/121665\">https://mathoverflow.net/a/297392/121665</a></p>\n" }, { "answer_id": 260634, "score": 27, "text": "<p>Unsurprisingly, the topics that occur to me have various connections to number theory (and related harmonic analysis) (and unclear to me what might have already been done in your course...):</p>\n\n<p>EDIT: inserted some links... EDIT-EDIT: one more...</p>\n\n<ul>\n<li><p>Genuinely distributional proof of Poisson summation: <a href=\"http://www.math.umn.edu/~garrett/m/fun/poisson.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/poisson.pdf</a></p></li>\n<li><p>Meromorphic continuation of distributions $|x|^s$ and ${\\mathrm sgn}(x)\\cdot |x|^s$ <a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf</a></p></li>\n<li><p>(Fancier version of the previous: mero cont'n of $|\\det x|^s$ on $n\\times n$ matrices (if this is interesting, I have some notes, or maybe it's a fun exercise). In particular, stimulated by a math-overflow question of A. Braverman some time ago, there <em>are</em> equivariant distributions (e.g., on two-by-two matrices) such that both they and their Fourier transform are supported on singular matrices... Wacky! <a href=\"http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf</a></p></li>\n<li><p>Decomposition of $L^2(A)$ for compact abelian topological groups $A$ (by Hilbert-Schmidt, hence compact, operators). <a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf</a></p></li>\n<li><p>Reconsideration of Sturm-Liouville problems (with reasonable hypotheses), to really <em>prove</em> things that are ... ahem... \"suggested\" in usual more-naive discussions.</p></li>\n<li><p>Quadratic reciprocity over number fields (and function fields not of char=2) ... cf. <a href=\"http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf</a>\n(This presumes Poisson summation for $\\mathbb A/k$...)</p></li>\n<li><p>Explanation that Schwartz' kernel theorem is a corollary of a Cartan-Eilenberg adjunction (between $\\otimes$ and $\\mathrm{Hom(,-)}$), when we know that there are genuine (i.e., categorically correct) tensor products for \"nuclear Frechet spaces\", ... which leads to the issue of suitable notions of the latter. <a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf</a></p></li>\n<li><p>The idea that termwise differentiation of Fourier series is \"<em>always</em> ok\" (with coefs that grow at most polynomially), if/when the outcome is interpreted as lying in a suitable Sobolev space on the circle. And that polynomial-growth-coefficiented Fourier series _always_converge_... if only in a suitable Sobolev space.\n<a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf</a></p></li>\n</ul>\n\n<p>** Possibility of ranting about the limitations of pointwise convergence... especially when placed in contrast to convergence in Sobolev spaces...</p>\n\n<ul>\n<li><p>Use of Snake Lemma to talk about mero cont'n of the Gamma function, via mero cont'n of $|x|^s$. :) <a href=\"http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf</a></p></li>\n<li><p>Peetre's theorem that any (not necessarily continuous!) linear functional on test functions that does not increase support is a differential operator. (I have a note on this, which may be more palatable to beginners than Peetre's paper.) <a href=\"http://www.math.umn.edu/~garrett/m/v/peetre.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/peetre.pdf</a></p></li>\n<li><p>Uniqueness of invariant functionals... As the easiest case (which is easy, but cognitive-dissonance-provoking, in my experience), proving that $u'=0$ for a distribution $u$ implies that $u$ is (integration-against) a constant. (Maybe you'd do this along the way...) <a href=\"http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf</a></p></li>\n<li><p>... this is not a stand-alone topic, but: the usual discussions of pseudo-differential operators (e.g., \"symbols\") somehow shrink from talking about <em>quotients</em> of TVS's in a grown-up way... If that hadn't been done earlier, and/or people had a (reasonable!) feeling of discomfort about the usual style of chatter in the psi-DO world, perhaps this could be happy-making.</p></li>\n<li><p>Meromorphic/holomorphic vector-valued functions (cf. Grothendieck c. 1953-4, and also various of my notes) e.g., meromorphic families of distributions... E.g., the $|x|^s$ family on $\\mathbb R$ has <em>residues</em> which are the even-order derivatives of $\\delta$, and ${\\mathrm {sgn}}(x)\\cdot |x|^s$ has as residues the odd-order derivatives of $\\delta$. <a href=\"http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf</a>\nDepending on context, there are somewhat-fancier things that I do find entertaining and also useful. Comments/correspondence are welcome.</p></li>\n</ul>\n" }, { "answer_id": 259874, "score": 12, "text": "<p>Two (<strong>edit:</strong> now four) not-so-usual examples come to my mind:</p>\n\n<ul>\n<li>There is the proof of the <em>central limit theorem</em> using Fourier analysis, as done in Chapter 7 of Hörmander's book. It's a cornerstone of probability theory and the Fourier-analytic proof presented by Hörmander is a nice showcase of the power of distribution theory: one gets a conceptually deep and useful theorem in full generality with relatively little effort;</li>\n<li>Something I have been meaning to try is to work out some \"mathematical toy examples\" of continuum limits of distributions supported on a lattice, as done e.g. in Chapter 15 of the book by R. Fernández, J. Fröhlich and A.D. Sokal, <em>Random Walks, Critical Phenomena, and Triviality in Quantum Field Theory</em> (Springer, 1992). I do not know if it is a feasible task, but since you are a specialist is constructive QFT, you should be able to do it... <strong>Edit 3:</strong> a classical and rather charming example of this has just came to my attention: the so-called <em>Bernstein polynomials</em>, used by their namesake Sergei N. Bernstein (the same mathematician who found the family of inequalities - also named after him - for smooth functions with compactly supported Fourier transform which underlie the Littlewood-Paley theory discussed in my Edit 1 below) to provide a probabilistic proof of the Weierstrass approximation theorem through the weak law of large numbers. Taking formal adjoints of the sequence of Bernstein polynomial operators provides a general lattice approximation for distributions on compact rectangular domains. A rather extensive discussion can be found in the short and beautiful book of the same title by G. G. Lorentz (the mathematician who invented the so-called Lorentz interpolation spaces, not the Dutch physicist).</li>\n<li><strong>(Edit 1)</strong> A third one, which I am particularly fond of, is the dyadic Fourier analysis of Littlewood and Paley, which is strongly connected to point (3). The connection to renormalization group ideas is quite obvious, and it has become a standard tool for obtaining important inequalities in nonlinear analysis (Gagliardo-Nirenberg, Moser, Schauder), all the way to the monumental Nash-Moser implicit function theorem (as shown by Hörmander - him again - in the 70's). A nice pedagogical exposition of these ideas can be found the book by S. Alinhac and P. Gérard, <em>Pseudo-Differential Operators and the Nash-Moser Theorem</em> (AMS, 2007) - just watch out for the numerous typographic errors!</li>\n<li><strong>(Edit 2)</strong> A fourth one, which is not actually unusual in itself but rather in its standpoint, is that many interesting examples of distributions (principal values, <em>partes finies</em>, extensions of homogeneous distributions to the origin, etc.) are obtained as extensions of distributions initially defined in (often finite-codimensional) subspaces of test functions, collectively called <em>renormalizations</em>. What would be interesting, also regarding applications to physics, is to treat such examples from this unified viewpoint. Indeed, as it can be seen e.g. by (Bogoliubov-Parasiuk-)Hepp's and Epstein-Glaser's treatments of perturbative renormalization in QFT, the latter really amounts to renormalizing distributions in the above sense. As a famous dictum of K. Hepp goes (<em>Proof of the Bogoliubov-Parasiuk Theorem on Renormalization</em>, Commun. Math. Phys. <strong>2</strong> (1966) 301-326):\n\n<blockquote>\n <p>\"The renormalization theory (of Dyson) is in this framework a constructive \n form of the Hahn-Banach theorem.\" (pp. 302 - parentheses mine)</p>\n</blockquote></li>\n</ul>\n" }, { "answer_id": 259864, "score": 6, "text": "<p>One of my favourite applications of 'basic' distribution theory, which actually requires most of the tools in your list to be fully apreciated, is the Malgrange--Ehrenpreis Theorem on the local solvability of arbitrary constant coefficients PDE. There are a few proofs, but the proof in M.Taylor's book on Pseudodifferential Operators (last Section of first Chapter) is especially suited for your course I think. It is a substantial theorem though, so it might or might be not a good idea to include it, depending on the length and size of the course.</p>\n" }, { "answer_id": 289039, "score": 3, "text": "<p>No so much as applications, but I would love to learn about the basics of hyperfunctions (of the Japanese school) in such a course, if only in one dimension. Specifically I think it gives an easier way to define the wavefront set, which is what's required to explain why we can't square the delta function (referring to a comment above). See also the question <a href=\"https://mathoverflow.net/questions/48067/is-square-of-delta-function-defined-somewhere\">Is square of Delta function defined somewhere?</a></p>\n\n<p>For what it's worth, I found the book</p>\n\n<ul>\n<li>Kato &amp; Struppa, <em>Fundamentals of Algebraic Microlocal Analysis</em></li>\n</ul>\n\n<p>particularly well written. The article in EoM is also great</p>\n\n<p><a href=\"https://www.encyclopediaofmath.org/index.php/Hyperfunction\" rel=\"nofollow noreferrer\">https://www.encyclopediaofmath.org/index.php/Hyperfunction</a></p>\n" }, { "answer_id": 298578, "score": 1, "text": "<p>Let me mention two nice applications to Sobolev spaces:</p>\n\n<p>The first result shows that if $f\\in \\mathcal{D}'$ and $\\nabla f\\in L^p_{\\rm loc}$, then $f\\in L^p_{\\rm loc}$. For a short proof of this classical result see\n<a href=\"https://mathoverflow.net/a/296464/121665\">https://mathoverflow.net/a/296464/121665</a></p>\n\n<p>For applications of this result to higher order Sobolev spaces see:<a href=\"https://mathoverflow.net/a/297392/121665\">https://mathoverflow.net/a/297392/121665</a></p>\n" }, { "answer_id": 260634, "score": 27, "text": "<p>Unsurprisingly, the topics that occur to me have various connections to number theory (and related harmonic analysis) (and unclear to me what might have already been done in your course...):</p>\n\n<p>EDIT: inserted some links... EDIT-EDIT: one more...</p>\n\n<ul>\n<li><p>Genuinely distributional proof of Poisson summation: <a href=\"http://www.math.umn.edu/~garrett/m/fun/poisson.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/poisson.pdf</a></p></li>\n<li><p>Meromorphic continuation of distributions $|x|^s$ and ${\\mathrm sgn}(x)\\cdot |x|^s$ <a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf</a></p></li>\n<li><p>(Fancier version of the previous: mero cont'n of $|\\det x|^s$ on $n\\times n$ matrices (if this is interesting, I have some notes, or maybe it's a fun exercise). In particular, stimulated by a math-overflow question of A. Braverman some time ago, there <em>are</em> equivariant distributions (e.g., on two-by-two matrices) such that both they and their Fourier transform are supported on singular matrices... Wacky! <a href=\"http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf</a></p></li>\n<li><p>Decomposition of $L^2(A)$ for compact abelian topological groups $A$ (by Hilbert-Schmidt, hence compact, operators). <a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf</a></p></li>\n<li><p>Reconsideration of Sturm-Liouville problems (with reasonable hypotheses), to really <em>prove</em> things that are ... ahem... \"suggested\" in usual more-naive discussions.</p></li>\n<li><p>Quadratic reciprocity over number fields (and function fields not of char=2) ... cf. <a href=\"http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf</a>\n(This presumes Poisson summation for $\\mathbb A/k$...)</p></li>\n<li><p>Explanation that Schwartz' kernel theorem is a corollary of a Cartan-Eilenberg adjunction (between $\\otimes$ and $\\mathrm{Hom(,-)}$), when we know that there are genuine (i.e., categorically correct) tensor products for \"nuclear Frechet spaces\", ... which leads to the issue of suitable notions of the latter. <a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf</a></p></li>\n<li><p>The idea that termwise differentiation of Fourier series is \"<em>always</em> ok\" (with coefs that grow at most polynomially), if/when the outcome is interpreted as lying in a suitable Sobolev space on the circle. And that polynomial-growth-coefficiented Fourier series _always_converge_... if only in a suitable Sobolev space.\n<a href=\"http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf</a></p></li>\n</ul>\n\n<p>** Possibility of ranting about the limitations of pointwise convergence... especially when placed in contrast to convergence in Sobolev spaces...</p>\n\n<ul>\n<li><p>Use of Snake Lemma to talk about mero cont'n of the Gamma function, via mero cont'n of $|x|^s$. :) <a href=\"http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf</a></p></li>\n<li><p>Peetre's theorem that any (not necessarily continuous!) linear functional on test functions that does not increase support is a differential operator. (I have a note on this, which may be more palatable to beginners than Peetre's paper.) <a href=\"http://www.math.umn.edu/~garrett/m/v/peetre.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/peetre.pdf</a></p></li>\n<li><p>Uniqueness of invariant functionals... As the easiest case (which is easy, but cognitive-dissonance-provoking, in my experience), proving that $u'=0$ for a distribution $u$ implies that $u$ is (integration-against) a constant. (Maybe you'd do this along the way...) <a href=\"http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf</a></p></li>\n<li><p>... this is not a stand-alone topic, but: the usual discussions of pseudo-differential operators (e.g., \"symbols\") somehow shrink from talking about <em>quotients</em> of TVS's in a grown-up way... If that hadn't been done earlier, and/or people had a (reasonable!) feeling of discomfort about the usual style of chatter in the psi-DO world, perhaps this could be happy-making.</p></li>\n<li><p>Meromorphic/holomorphic vector-valued functions (cf. Grothendieck c. 1953-4, and also various of my notes) e.g., meromorphic families of distributions... E.g., the $|x|^s$ family on $\\mathbb R$ has <em>residues</em> which are the even-order derivatives of $\\delta$, and ${\\mathrm {sgn}}(x)\\cdot |x|^s$ has as residues the odd-order derivatives of $\\delta$. <a href=\"http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf\" rel=\"noreferrer\">http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf</a>\nDepending on context, there are somewhat-fancier things that I do find entertaining and also useful. Comments/correspondence are welcome.</p></li>\n</ul>\n" }, { "answer_id": 259874, "score": 12, "text": "<p>Two (<strong>edit:</strong> now four) not-so-usual examples come to my mind:</p>\n\n<ul>\n<li>There is the proof of the <em>central limit theorem</em> using Fourier analysis, as done in Chapter 7 of Hörmander's book. It's a cornerstone of probability theory and the Fourier-analytic proof presented by Hörmander is a nice showcase of the power of distribution theory: one gets a conceptually deep and useful theorem in full generality with relatively little effort;</li>\n<li>Something I have been meaning to try is to work out some \"mathematical toy examples\" of continuum limits of distributions supported on a lattice, as done e.g. in Chapter 15 of the book by R. Fernández, J. Fröhlich and A.D. Sokal, <em>Random Walks, Critical Phenomena, and Triviality in Quantum Field Theory</em> (Springer, 1992). I do not know if it is a feasible task, but since you are a specialist is constructive QFT, you should be able to do it... <strong>Edit 3:</strong> a classical and rather charming example of this has just came to my attention: the so-called <em>Bernstein polynomials</em>, used by their namesake Sergei N. Bernstein (the same mathematician who found the family of inequalities - also named after him - for smooth functions with compactly supported Fourier transform which underlie the Littlewood-Paley theory discussed in my Edit 1 below) to provide a probabilistic proof of the Weierstrass approximation theorem through the weak law of large numbers. Taking formal adjoints of the sequence of Bernstein polynomial operators provides a general lattice approximation for distributions on compact rectangular domains. A rather extensive discussion can be found in the short and beautiful book of the same title by G. G. Lorentz (the mathematician who invented the so-called Lorentz interpolation spaces, not the Dutch physicist).</li>\n<li><strong>(Edit 1)</strong> A third one, which I am particularly fond of, is the dyadic Fourier analysis of Littlewood and Paley, which is strongly connected to point (3). The connection to renormalization group ideas is quite obvious, and it has become a standard tool for obtaining important inequalities in nonlinear analysis (Gagliardo-Nirenberg, Moser, Schauder), all the way to the monumental Nash-Moser implicit function theorem (as shown by Hörmander - him again - in the 70's). A nice pedagogical exposition of these ideas can be found the book by S. Alinhac and P. Gérard, <em>Pseudo-Differential Operators and the Nash-Moser Theorem</em> (AMS, 2007) - just watch out for the numerous typographic errors!</li>\n<li><strong>(Edit 2)</strong> A fourth one, which is not actually unusual in itself but rather in its standpoint, is that many interesting examples of distributions (principal values, <em>partes finies</em>, extensions of homogeneous distributions to the origin, etc.) are obtained as extensions of distributions initially defined in (often finite-codimensional) subspaces of test functions, collectively called <em>renormalizations</em>. What would be interesting, also regarding applications to physics, is to treat such examples from this unified viewpoint. Indeed, as it can be seen e.g. by (Bogoliubov-Parasiuk-)Hepp's and Epstein-Glaser's treatments of perturbative renormalization in QFT, the latter really amounts to renormalizing distributions in the above sense. As a famous dictum of K. Hepp goes (<em>Proof of the Bogoliubov-Parasiuk Theorem on Renormalization</em>, Commun. Math. Phys. <strong>2</strong> (1966) 301-326):\n\n<blockquote>\n <p>\"The renormalization theory (of Dyson) is in this framework a constructive \n form of the Hahn-Banach theorem.\" (pp. 302 - parentheses mine)</p>\n</blockquote></li>\n</ul>\n" }, { "answer_id": 259864, "score": 6, "text": "<p>One of my favourite applications of 'basic' distribution theory, which actually requires most of the tools in your list to be fully apreciated, is the Malgrange--Ehrenpreis Theorem on the local solvability of arbitrary constant coefficients PDE. There are a few proofs, but the proof in M.Taylor's book on Pseudodifferential Operators (last Section of first Chapter) is especially suited for your course I think. It is a substantial theorem though, so it might or might be not a good idea to include it, depending on the length and size of the course.</p>\n" }, { "answer_id": 289039, "score": 3, "text": "<p>No so much as applications, but I would love to learn about the basics of hyperfunctions (of the Japanese school) in such a course, if only in one dimension. Specifically I think it gives an easier way to define the wavefront set, which is what's required to explain why we can't square the delta function (referring to a comment above). See also the question <a href=\"https://mathoverflow.net/questions/48067/is-square-of-delta-function-defined-somewhere\">Is square of Delta function defined somewhere?</a></p>\n\n<p>For what it's worth, I found the book</p>\n\n<ul>\n<li>Kato &amp; Struppa, <em>Fundamentals of Algebraic Microlocal Analysis</em></li>\n</ul>\n\n<p>particularly well written. The article in EoM is also great</p>\n\n<p><a href=\"https://www.encyclopediaofmath.org/index.php/Hyperfunction\" rel=\"nofollow noreferrer\">https://www.encyclopediaofmath.org/index.php/Hyperfunction</a></p>\n" }, { "answer_id": 298578, "score": 1, "text": "<p>Let me mention two nice applications to Sobolev spaces:</p>\n\n<p>The first result shows that if $f\\in \\mathcal{D}'$ and $\\nabla f\\in L^p_{\\rm loc}$, then $f\\in L^p_{\\rm loc}$. For a short proof of this classical result see\n<a href=\"https://mathoverflow.net/a/296464/121665\">https://mathoverflow.net/a/296464/121665</a></p>\n\n<p>For applications of this result to higher order Sobolev spaces see:<a href=\"https://mathoverflow.net/a/297392/121665\">https://mathoverflow.net/a/297392/121665</a></p>\n" } ]
Question: Nice applications for Schwartz distributions Answer 1: <p>Unsurprisingly, the topics that occur to me have various connections to number theory (and related harmonic analysis) (and unclear to me what might have already been done in your course...):</p> <p>EDIT: inserted some links... EDIT-EDIT: one more...</p> <ul> <li><p>Genuinely distributional proof of Poisson summation: <a href="http://www.math.umn.edu/~garrett/m/fun/poisson.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/poisson.pdf</a></p></li> <li><p>Meromorphic continuation of distributions $|x|^s$ and ${\mathrm sgn}(x)\cdot |x|^s$ <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf</a></p></li> <li><p>(Fancier version of the previous: mero cont'n of $|\det x|^s$ on $n\times n$ matrices (if this is interesting, I have some notes, or maybe it's a fun exercise). In particular, stimulated by a math-overflow question of A. Braverman some time ago, there <em>are</em> equivariant distributions (e.g., on two-by-two matrices) such that both they and their Fourier transform are supported on singular matrices... Wacky! <a href="http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf</a></p></li> <li><p>Decomposition of $L^2(A)$ for compact abelian topological groups $A$ (by Hilbert-Schmidt, hence compact, operators). <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf</a></p></li> <li><p>Reconsideration of Sturm-Liouville problems (with reasonable hypotheses), to really <em>prove</em> things that are ... ahem... "suggested" in usual more-naive discussions.</p></li> <li><p>Quadratic reciprocity over number fields (and function fields not of char=2) ... cf. <a href="http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf</a> (This presumes Poisson summation for $\mathbb A/k$...)</p></li> <li><p>Explanation that Schwartz' kernel theorem is a corollary of a Cartan-Eilenberg adjunction (between $\otimes$ and $\mathrm{Hom(,-)}$), when we know that there are genuine (i.e., categorically correct) tensor products for "nuclear Frechet spaces", ... which leads to the issue of suitable notions of the latter. <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf</a></p></li> <li><p>The idea that termwise differentiation of Fourier series is "<em>always</em> ok" (with coefs that grow at most polynomially), if/when the outcome is interpreted as lying in a suitable Sobolev space on the circle. And that polynomial-growth-coefficiented Fourier series _always_converge_... if only in a suitable Sobolev space. <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf</a></p></li> </ul> <p>** Possibility of ranting about the limitations of pointwise convergence... especially when placed in contrast to convergence in Sobolev spaces...</p> <ul> <li><p>Use of Snake Lemma to talk about mero cont'n of the Gamma function, via mero cont'n of $|x|^s$. :) <a href="http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf</a></p></li> <li><p>Peetre's theorem that any (not necessarily continuous!) linear functional on test functions that does not increase support is a differential operator. (I have a note on this, which may be more palatable to beginners than Peetre's paper.) <a href="http://www.math.umn.edu/~garrett/m/v/peetre.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/peetre.pdf</a></p></li> <li><p>Uniqueness of invariant functionals... As the easiest case (which is easy, but cognitive-dissonance-provoking, in my experience), proving that $u'=0$ for a distribution $u$ implies that $u$ is (integration-against) a constant. (Maybe you'd do this along the way...) <a href="http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf</a></p></li> <li><p>... this is not a stand-alone topic, but: the usual discussions of pseudo-differential operators (e.g., "symbols") somehow shrink from talking about <em>quotients</em> of TVS's in a grown-up way... If that hadn't been done earlier, and/or people had a (reasonable!) feeling of discomfort about the usual style of chatter in the psi-DO world, perhaps this could be happy-making.</p></li> <li><p>Meromorphic/holomorphic vector-valued functions (cf. Grothendieck c. 1953-4, and also various of my notes) e.g., meromorphic families of distributions... E.g., the $|x|^s$ family on $\mathbb R$ has <em>residues</em> which are the even-order derivatives of $\delta$, and ${\mathrm {sgn}}(x)\cdot |x|^s$ has as residues the odd-order derivatives of $\delta$. <a href="http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf</a> Depending on context, there are somewhat-fancier things that I do find entertaining and also useful. Comments/correspondence are welcome.</p></li> </ul> Answer 2: <p>Two (<strong>edit:</strong> now four) not-so-usual examples come to my mind:</p> <ul> <li>There is the proof of the <em>central limit theorem</em> using Fourier analysis, as done in Chapter 7 of Hörmander's book. It's a cornerstone of probability theory and the Fourier-analytic proof presented by Hörmander is a nice showcase of the power of distribution theory: one gets a conceptually deep and useful theorem in full generality with relatively little effort;</li> <li>Something I have been meaning to try is to work out some "mathematical toy examples" of continuum limits of distributions supported on a lattice, as done e.g. in Chapter 15 of the book by R. Fernández, J. Fröhlich and A.D. Sokal, <em>Random Walks, Critical Phenomena, and Triviality in Quantum Field Theory</em> (Springer, 1992). I do not know if it is a feasible task, but since you are a specialist is constructive QFT, you should be able to do it... <strong>Edit 3:</strong> a classical and rather charming example of this has just came to my attention: the so-called <em>Bernstein polynomials</em>, used by their namesake Sergei N. Bernstein (the same mathematician who found the family of inequalities - also named after him - for smooth functions with compactly supported Fourier transform which underlie the Littlewood-Paley theory discussed in my Edit 1 below) to provide a probabilistic proof of the Weierstrass approximation theorem through the weak law of large numbers. Taking formal adjoints of the sequence of Bernstein polynomial operators provides a general lattice approximation for distributions on compact rectangular domains. A rather extensive discussion can be found in the short and beautiful book of the same title by G. G. Lorentz (the mathematician who invented the so-called Lorentz interpolation spaces, not the Dutch physicist).</li> <li><strong>(Edit 1)</strong> A third one, which I am particularly fond of, is the dyadic Fourier analysis of Littlewood and Paley, which is strongly connected to point (3). The connection to renormalization group ideas is quite obvious, and it has become a standard tool for obtaining important inequalities in nonlinear analysis (Gagliardo-Nirenberg, Moser, Schauder), all the way to the monumental Nash-Moser implicit function theorem (as shown by Hörmander - him again - in the 70's). A nice pedagogical exposition of these ideas can be found the book by S. Alinhac and P. Gérard, <em>Pseudo-Differential Operators and the Nash-Moser Theorem</em> (AMS, 2007) - just watch out for the numerous typographic errors!</li> <li><strong>(Edit 2)</strong> A fourth one, which is not actually unusual in itself but rather in its standpoint, is that many interesting examples of distributions (principal values, <em>partes finies</em>, extensions of homogeneous distributions to the origin, etc.) are obtained as extensions of distributions initially defined in (often finite-codimensional) subspaces of test functions, collectively called <em>renormalizations</em>. What would be interesting, also regarding applications to physics, is to treat such examples from this unified viewpoint. Indeed, as it can be seen e.g. by (Bogoliubov-Parasiuk-)Hepp's and Epstein-Glaser's treatments of perturbative renormalization in QFT, the latter really amounts to renormalizing distributions in the above sense. As a famous dictum of K. Hepp goes (<em>Proof of the Bogoliubov-Parasiuk Theorem on Renormalization</em>, Commun. Math. Phys. <strong>2</strong> (1966) 301-326): <blockquote> <p>"The renormalization theory (of Dyson) is in this framework a constructive form of the Hahn-Banach theorem." (pp. 302 - parentheses mine)</p> </blockquote></li> </ul> Answer 3: <p>One of my favourite applications of 'basic' distribution theory, which actually requires most of the tools in your list to be fully apreciated, is the Malgrange--Ehrenpreis Theorem on the local solvability of arbitrary constant coefficients PDE. There are a few proofs, but the proof in M.Taylor's book on Pseudodifferential Operators (last Section of first Chapter) is especially suited for your course I think. It is a substantial theorem though, so it might or might be not a good idea to include it, depending on the length and size of the course.</p> Answer 4: <p>No so much as applications, but I would love to learn about the basics of hyperfunctions (of the Japanese school) in such a course, if only in one dimension. Specifically I think it gives an easier way to define the wavefront set, which is what's required to explain why we can't square the delta function (referring to a comment above). See also the question <a href="https://mathoverflow.net/questions/48067/is-square-of-delta-function-defined-somewhere">Is square of Delta function defined somewhere?</a></p> <p>For what it's worth, I found the book</p> <ul> <li>Kato &amp; Struppa, <em>Fundamentals of Algebraic Microlocal Analysis</em></li> </ul> <p>particularly well written. The article in EoM is also great</p> <p><a href="https://www.encyclopediaofmath.org/index.php/Hyperfunction" rel="nofollow noreferrer">https://www.encyclopediaofmath.org/index.php/Hyperfunction</a></p> Answer 5: <p>Let me mention two nice applications to Sobolev spaces:</p> <p>The first result shows that if $f\in \mathcal{D}'$ and $\nabla f\in L^p_{\rm loc}$, then $f\in L^p_{\rm loc}$. For a short proof of this classical result see <a href="https://mathoverflow.net/a/296464/121665">https://mathoverflow.net/a/296464/121665</a></p> <p>For applications of this result to higher order Sobolev spaces see:<a href="https://mathoverflow.net/a/297392/121665">https://mathoverflow.net/a/297392/121665</a></p> Answer 6: <p>Unsurprisingly, the topics that occur to me have various connections to number theory (and related harmonic analysis) (and unclear to me what might have already been done in your course...):</p> <p>EDIT: inserted some links... EDIT-EDIT: one more...</p> <ul> <li><p>Genuinely distributional proof of Poisson summation: <a href="http://www.math.umn.edu/~garrett/m/fun/poisson.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/poisson.pdf</a></p></li> <li><p>Meromorphic continuation of distributions $|x|^s$ and ${\mathrm sgn}(x)\cdot |x|^s$ <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf</a></p></li> <li><p>(Fancier version of the previous: mero cont'n of $|\det x|^s$ on $n\times n$ matrices (if this is interesting, I have some notes, or maybe it's a fun exercise). In particular, stimulated by a math-overflow question of A. Braverman some time ago, there <em>are</em> equivariant distributions (e.g., on two-by-two matrices) such that both they and their Fourier transform are supported on singular matrices... Wacky! <a href="http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf</a></p></li> <li><p>Decomposition of $L^2(A)$ for compact abelian topological groups $A$ (by Hilbert-Schmidt, hence compact, operators). <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf</a></p></li> <li><p>Reconsideration of Sturm-Liouville problems (with reasonable hypotheses), to really <em>prove</em> things that are ... ahem... "suggested" in usual more-naive discussions.</p></li> <li><p>Quadratic reciprocity over number fields (and function fields not of char=2) ... cf. <a href="http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf</a> (This presumes Poisson summation for $\mathbb A/k$...)</p></li> <li><p>Explanation that Schwartz' kernel theorem is a corollary of a Cartan-Eilenberg adjunction (between $\otimes$ and $\mathrm{Hom(,-)}$), when we know that there are genuine (i.e., categorically correct) tensor products for "nuclear Frechet spaces", ... which leads to the issue of suitable notions of the latter. <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf</a></p></li> <li><p>The idea that termwise differentiation of Fourier series is "<em>always</em> ok" (with coefs that grow at most polynomially), if/when the outcome is interpreted as lying in a suitable Sobolev space on the circle. And that polynomial-growth-coefficiented Fourier series _always_converge_... if only in a suitable Sobolev space. <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf</a></p></li> </ul> <p>** Possibility of ranting about the limitations of pointwise convergence... especially when placed in contrast to convergence in Sobolev spaces...</p> <ul> <li><p>Use of Snake Lemma to talk about mero cont'n of the Gamma function, via mero cont'n of $|x|^s$. :) <a href="http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf</a></p></li> <li><p>Peetre's theorem that any (not necessarily continuous!) linear functional on test functions that does not increase support is a differential operator. (I have a note on this, which may be more palatable to beginners than Peetre's paper.) <a href="http://www.math.umn.edu/~garrett/m/v/peetre.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/peetre.pdf</a></p></li> <li><p>Uniqueness of invariant functionals... As the easiest case (which is easy, but cognitive-dissonance-provoking, in my experience), proving that $u'=0$ for a distribution $u$ implies that $u$ is (integration-against) a constant. (Maybe you'd do this along the way...) <a href="http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf</a></p></li> <li><p>... this is not a stand-alone topic, but: the usual discussions of pseudo-differential operators (e.g., "symbols") somehow shrink from talking about <em>quotients</em> of TVS's in a grown-up way... If that hadn't been done earlier, and/or people had a (reasonable!) feeling of discomfort about the usual style of chatter in the psi-DO world, perhaps this could be happy-making.</p></li> <li><p>Meromorphic/holomorphic vector-valued functions (cf. Grothendieck c. 1953-4, and also various of my notes) e.g., meromorphic families of distributions... E.g., the $|x|^s$ family on $\mathbb R$ has <em>residues</em> which are the even-order derivatives of $\delta$, and ${\mathrm {sgn}}(x)\cdot |x|^s$ has as residues the odd-order derivatives of $\delta$. <a href="http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf</a> Depending on context, there are somewhat-fancier things that I do find entertaining and also useful. Comments/correspondence are welcome.</p></li> </ul> Answer 7: <p>Two (<strong>edit:</strong> now four) not-so-usual examples come to my mind:</p> <ul> <li>There is the proof of the <em>central limit theorem</em> using Fourier analysis, as done in Chapter 7 of Hörmander's book. It's a cornerstone of probability theory and the Fourier-analytic proof presented by Hörmander is a nice showcase of the power of distribution theory: one gets a conceptually deep and useful theorem in full generality with relatively little effort;</li> <li>Something I have been meaning to try is to work out some "mathematical toy examples" of continuum limits of distributions supported on a lattice, as done e.g. in Chapter 15 of the book by R. Fernández, J. Fröhlich and A.D. Sokal, <em>Random Walks, Critical Phenomena, and Triviality in Quantum Field Theory</em> (Springer, 1992). I do not know if it is a feasible task, but since you are a specialist is constructive QFT, you should be able to do it... <strong>Edit 3:</strong> a classical and rather charming example of this has just came to my attention: the so-called <em>Bernstein polynomials</em>, used by their namesake Sergei N. Bernstein (the same mathematician who found the family of inequalities - also named after him - for smooth functions with compactly supported Fourier transform which underlie the Littlewood-Paley theory discussed in my Edit 1 below) to provide a probabilistic proof of the Weierstrass approximation theorem through the weak law of large numbers. Taking formal adjoints of the sequence of Bernstein polynomial operators provides a general lattice approximation for distributions on compact rectangular domains. A rather extensive discussion can be found in the short and beautiful book of the same title by G. G. Lorentz (the mathematician who invented the so-called Lorentz interpolation spaces, not the Dutch physicist).</li> <li><strong>(Edit 1)</strong> A third one, which I am particularly fond of, is the dyadic Fourier analysis of Littlewood and Paley, which is strongly connected to point (3). The connection to renormalization group ideas is quite obvious, and it has become a standard tool for obtaining important inequalities in nonlinear analysis (Gagliardo-Nirenberg, Moser, Schauder), all the way to the monumental Nash-Moser implicit function theorem (as shown by Hörmander - him again - in the 70's). A nice pedagogical exposition of these ideas can be found the book by S. Alinhac and P. Gérard, <em>Pseudo-Differential Operators and the Nash-Moser Theorem</em> (AMS, 2007) - just watch out for the numerous typographic errors!</li> <li><strong>(Edit 2)</strong> A fourth one, which is not actually unusual in itself but rather in its standpoint, is that many interesting examples of distributions (principal values, <em>partes finies</em>, extensions of homogeneous distributions to the origin, etc.) are obtained as extensions of distributions initially defined in (often finite-codimensional) subspaces of test functions, collectively called <em>renormalizations</em>. What would be interesting, also regarding applications to physics, is to treat such examples from this unified viewpoint. Indeed, as it can be seen e.g. by (Bogoliubov-Parasiuk-)Hepp's and Epstein-Glaser's treatments of perturbative renormalization in QFT, the latter really amounts to renormalizing distributions in the above sense. As a famous dictum of K. Hepp goes (<em>Proof of the Bogoliubov-Parasiuk Theorem on Renormalization</em>, Commun. Math. Phys. <strong>2</strong> (1966) 301-326): <blockquote> <p>"The renormalization theory (of Dyson) is in this framework a constructive form of the Hahn-Banach theorem." (pp. 302 - parentheses mine)</p> </blockquote></li> </ul> Answer 8: <p>One of my favourite applications of 'basic' distribution theory, which actually requires most of the tools in your list to be fully apreciated, is the Malgrange--Ehrenpreis Theorem on the local solvability of arbitrary constant coefficients PDE. There are a few proofs, but the proof in M.Taylor's book on Pseudodifferential Operators (last Section of first Chapter) is especially suited for your course I think. It is a substantial theorem though, so it might or might be not a good idea to include it, depending on the length and size of the course.</p> Answer 9: <p>No so much as applications, but I would love to learn about the basics of hyperfunctions (of the Japanese school) in such a course, if only in one dimension. Specifically I think it gives an easier way to define the wavefront set, which is what's required to explain why we can't square the delta function (referring to a comment above). See also the question <a href="https://mathoverflow.net/questions/48067/is-square-of-delta-function-defined-somewhere">Is square of Delta function defined somewhere?</a></p> <p>For what it's worth, I found the book</p> <ul> <li>Kato &amp; Struppa, <em>Fundamentals of Algebraic Microlocal Analysis</em></li> </ul> <p>particularly well written. The article in EoM is also great</p> <p><a href="https://www.encyclopediaofmath.org/index.php/Hyperfunction" rel="nofollow noreferrer">https://www.encyclopediaofmath.org/index.php/Hyperfunction</a></p> Answer 10: <p>Let me mention two nice applications to Sobolev spaces:</p> <p>The first result shows that if $f\in \mathcal{D}'$ and $\nabla f\in L^p_{\rm loc}$, then $f\in L^p_{\rm loc}$. For a short proof of this classical result see <a href="https://mathoverflow.net/a/296464/121665">https://mathoverflow.net/a/296464/121665</a></p> <p>For applications of this result to higher order Sobolev spaces see:<a href="https://mathoverflow.net/a/297392/121665">https://mathoverflow.net/a/297392/121665</a></p> Answer 11: <p>Unsurprisingly, the topics that occur to me have various connections to number theory (and related harmonic analysis) (and unclear to me what might have already been done in your course...):</p> <p>EDIT: inserted some links... EDIT-EDIT: one more...</p> <ul> <li><p>Genuinely distributional proof of Poisson summation: <a href="http://www.math.umn.edu/~garrett/m/fun/poisson.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/poisson.pdf</a></p></li> <li><p>Meromorphic continuation of distributions $|x|^s$ and ${\mathrm sgn}(x)\cdot |x|^s$ <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf</a></p></li> <li><p>(Fancier version of the previous: mero cont'n of $|\det x|^s$ on $n\times n$ matrices (if this is interesting, I have some notes, or maybe it's a fun exercise). In particular, stimulated by a math-overflow question of A. Braverman some time ago, there <em>are</em> equivariant distributions (e.g., on two-by-two matrices) such that both they and their Fourier transform are supported on singular matrices... Wacky! <a href="http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf</a></p></li> <li><p>Decomposition of $L^2(A)$ for compact abelian topological groups $A$ (by Hilbert-Schmidt, hence compact, operators). <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf</a></p></li> <li><p>Reconsideration of Sturm-Liouville problems (with reasonable hypotheses), to really <em>prove</em> things that are ... ahem... "suggested" in usual more-naive discussions.</p></li> <li><p>Quadratic reciprocity over number fields (and function fields not of char=2) ... cf. <a href="http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf</a> (This presumes Poisson summation for $\mathbb A/k$...)</p></li> <li><p>Explanation that Schwartz' kernel theorem is a corollary of a Cartan-Eilenberg adjunction (between $\otimes$ and $\mathrm{Hom(,-)}$), when we know that there are genuine (i.e., categorically correct) tensor products for "nuclear Frechet spaces", ... which leads to the issue of suitable notions of the latter. <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf</a></p></li> <li><p>The idea that termwise differentiation of Fourier series is "<em>always</em> ok" (with coefs that grow at most polynomially), if/when the outcome is interpreted as lying in a suitable Sobolev space on the circle. And that polynomial-growth-coefficiented Fourier series _always_converge_... if only in a suitable Sobolev space. <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf</a></p></li> </ul> <p>** Possibility of ranting about the limitations of pointwise convergence... especially when placed in contrast to convergence in Sobolev spaces...</p> <ul> <li><p>Use of Snake Lemma to talk about mero cont'n of the Gamma function, via mero cont'n of $|x|^s$. :) <a href="http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf</a></p></li> <li><p>Peetre's theorem that any (not necessarily continuous!) linear functional on test functions that does not increase support is a differential operator. (I have a note on this, which may be more palatable to beginners than Peetre's paper.) <a href="http://www.math.umn.edu/~garrett/m/v/peetre.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/peetre.pdf</a></p></li> <li><p>Uniqueness of invariant functionals... As the easiest case (which is easy, but cognitive-dissonance-provoking, in my experience), proving that $u'=0$ for a distribution $u$ implies that $u$ is (integration-against) a constant. (Maybe you'd do this along the way...) <a href="http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf</a></p></li> <li><p>... this is not a stand-alone topic, but: the usual discussions of pseudo-differential operators (e.g., "symbols") somehow shrink from talking about <em>quotients</em> of TVS's in a grown-up way... If that hadn't been done earlier, and/or people had a (reasonable!) feeling of discomfort about the usual style of chatter in the psi-DO world, perhaps this could be happy-making.</p></li> <li><p>Meromorphic/holomorphic vector-valued functions (cf. Grothendieck c. 1953-4, and also various of my notes) e.g., meromorphic families of distributions... E.g., the $|x|^s$ family on $\mathbb R$ has <em>residues</em> which are the even-order derivatives of $\delta$, and ${\mathrm {sgn}}(x)\cdot |x|^s$ has as residues the odd-order derivatives of $\delta$. <a href="http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf</a> Depending on context, there are somewhat-fancier things that I do find entertaining and also useful. Comments/correspondence are welcome.</p></li> </ul> Answer 12: <p>Two (<strong>edit:</strong> now four) not-so-usual examples come to my mind:</p> <ul> <li>There is the proof of the <em>central limit theorem</em> using Fourier analysis, as done in Chapter 7 of Hörmander's book. It's a cornerstone of probability theory and the Fourier-analytic proof presented by Hörmander is a nice showcase of the power of distribution theory: one gets a conceptually deep and useful theorem in full generality with relatively little effort;</li> <li>Something I have been meaning to try is to work out some "mathematical toy examples" of continuum limits of distributions supported on a lattice, as done e.g. in Chapter 15 of the book by R. Fernández, J. Fröhlich and A.D. Sokal, <em>Random Walks, Critical Phenomena, and Triviality in Quantum Field Theory</em> (Springer, 1992). I do not know if it is a feasible task, but since you are a specialist is constructive QFT, you should be able to do it... <strong>Edit 3:</strong> a classical and rather charming example of this has just came to my attention: the so-called <em>Bernstein polynomials</em>, used by their namesake Sergei N. Bernstein (the same mathematician who found the family of inequalities - also named after him - for smooth functions with compactly supported Fourier transform which underlie the Littlewood-Paley theory discussed in my Edit 1 below) to provide a probabilistic proof of the Weierstrass approximation theorem through the weak law of large numbers. Taking formal adjoints of the sequence of Bernstein polynomial operators provides a general lattice approximation for distributions on compact rectangular domains. A rather extensive discussion can be found in the short and beautiful book of the same title by G. G. Lorentz (the mathematician who invented the so-called Lorentz interpolation spaces, not the Dutch physicist).</li> <li><strong>(Edit 1)</strong> A third one, which I am particularly fond of, is the dyadic Fourier analysis of Littlewood and Paley, which is strongly connected to point (3). The connection to renormalization group ideas is quite obvious, and it has become a standard tool for obtaining important inequalities in nonlinear analysis (Gagliardo-Nirenberg, Moser, Schauder), all the way to the monumental Nash-Moser implicit function theorem (as shown by Hörmander - him again - in the 70's). A nice pedagogical exposition of these ideas can be found the book by S. Alinhac and P. Gérard, <em>Pseudo-Differential Operators and the Nash-Moser Theorem</em> (AMS, 2007) - just watch out for the numerous typographic errors!</li> <li><strong>(Edit 2)</strong> A fourth one, which is not actually unusual in itself but rather in its standpoint, is that many interesting examples of distributions (principal values, <em>partes finies</em>, extensions of homogeneous distributions to the origin, etc.) are obtained as extensions of distributions initially defined in (often finite-codimensional) subspaces of test functions, collectively called <em>renormalizations</em>. What would be interesting, also regarding applications to physics, is to treat such examples from this unified viewpoint. Indeed, as it can be seen e.g. by (Bogoliubov-Parasiuk-)Hepp's and Epstein-Glaser's treatments of perturbative renormalization in QFT, the latter really amounts to renormalizing distributions in the above sense. As a famous dictum of K. Hepp goes (<em>Proof of the Bogoliubov-Parasiuk Theorem on Renormalization</em>, Commun. Math. Phys. <strong>2</strong> (1966) 301-326): <blockquote> <p>"The renormalization theory (of Dyson) is in this framework a constructive form of the Hahn-Banach theorem." (pp. 302 - parentheses mine)</p> </blockquote></li> </ul> Answer 13: <p>One of my favourite applications of 'basic' distribution theory, which actually requires most of the tools in your list to be fully apreciated, is the Malgrange--Ehrenpreis Theorem on the local solvability of arbitrary constant coefficients PDE. There are a few proofs, but the proof in M.Taylor's book on Pseudodifferential Operators (last Section of first Chapter) is especially suited for your course I think. It is a substantial theorem though, so it might or might be not a good idea to include it, depending on the length and size of the course.</p> Answer 14: <p>No so much as applications, but I would love to learn about the basics of hyperfunctions (of the Japanese school) in such a course, if only in one dimension. Specifically I think it gives an easier way to define the wavefront set, which is what's required to explain why we can't square the delta function (referring to a comment above). See also the question <a href="https://mathoverflow.net/questions/48067/is-square-of-delta-function-defined-somewhere">Is square of Delta function defined somewhere?</a></p> <p>For what it's worth, I found the book</p> <ul> <li>Kato &amp; Struppa, <em>Fundamentals of Algebraic Microlocal Analysis</em></li> </ul> <p>particularly well written. The article in EoM is also great</p> <p><a href="https://www.encyclopediaofmath.org/index.php/Hyperfunction" rel="nofollow noreferrer">https://www.encyclopediaofmath.org/index.php/Hyperfunction</a></p> Answer 15: <p>Let me mention two nice applications to Sobolev spaces:</p> <p>The first result shows that if $f\in \mathcal{D}'$ and $\nabla f\in L^p_{\rm loc}$, then $f\in L^p_{\rm loc}$. For a short proof of this classical result see <a href="https://mathoverflow.net/a/296464/121665">https://mathoverflow.net/a/296464/121665</a></p> <p>For applications of this result to higher order Sobolev spaces see:<a href="https://mathoverflow.net/a/297392/121665">https://mathoverflow.net/a/297392/121665</a></p> Answer 16: <p>Unsurprisingly, the topics that occur to me have various connections to number theory (and related harmonic analysis) (and unclear to me what might have already been done in your course...):</p> <p>EDIT: inserted some links... EDIT-EDIT: one more...</p> <ul> <li><p>Genuinely distributional proof of Poisson summation: <a href="http://www.math.umn.edu/~garrett/m/fun/poisson.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/poisson.pdf</a></p></li> <li><p>Meromorphic continuation of distributions $|x|^s$ and ${\mathrm sgn}(x)\cdot |x|^s$ <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2013-14/mero_contn.pdf</a></p></li> <li><p>(Fancier version of the previous: mero cont'n of $|\det x|^s$ on $n\times n$ matrices (if this is interesting, I have some notes, or maybe it's a fun exercise). In particular, stimulated by a math-overflow question of A. Braverman some time ago, there <em>are</em> equivariant distributions (e.g., on two-by-two matrices) such that both they and their Fourier transform are supported on singular matrices... Wacky! <a href="http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/det_power_distn.pdf</a></p></li> <li><p>Decomposition of $L^2(A)$ for compact abelian topological groups $A$ (by Hilbert-Schmidt, hence compact, operators). <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06c_cpt_ab_gps.pdf</a></p></li> <li><p>Reconsideration of Sturm-Liouville problems (with reasonable hypotheses), to really <em>prove</em> things that are ... ahem... "suggested" in usual more-naive discussions.</p></li> <li><p>Quadratic reciprocity over number fields (and function fields not of char=2) ... cf. <a href="http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/quad_rec_02.pdf</a> (This presumes Poisson summation for $\mathbb A/k$...)</p></li> <li><p>Explanation that Schwartz' kernel theorem is a corollary of a Cartan-Eilenberg adjunction (between $\otimes$ and $\mathrm{Hom(,-)}$), when we know that there are genuine (i.e., categorically correct) tensor products for "nuclear Frechet spaces", ... which leads to the issue of suitable notions of the latter. <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/06d_nuclear_spaces_I.pdf</a></p></li> <li><p>The idea that termwise differentiation of Fourier series is "<em>always</em> ok" (with coefs that grow at most polynomially), if/when the outcome is interpreted as lying in a suitable Sobolev space on the circle. And that polynomial-growth-coefficiented Fourier series _always_converge_... if only in a suitable Sobolev space. <a href="http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/notes_2012-13/03b_intro_blevi.pdf</a></p></li> </ul> <p>** Possibility of ranting about the limitations of pointwise convergence... especially when placed in contrast to convergence in Sobolev spaces...</p> <ul> <li><p>Use of Snake Lemma to talk about mero cont'n of the Gamma function, via mero cont'n of $|x|^s$. :) <a href="http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/snake_lemma_gamma.pdf</a></p></li> <li><p>Peetre's theorem that any (not necessarily continuous!) linear functional on test functions that does not increase support is a differential operator. (I have a note on this, which may be more palatable to beginners than Peetre's paper.) <a href="http://www.math.umn.edu/~garrett/m/v/peetre.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/peetre.pdf</a></p></li> <li><p>Uniqueness of invariant functionals... As the easiest case (which is easy, but cognitive-dissonance-provoking, in my experience), proving that $u'=0$ for a distribution $u$ implies that $u$ is (integration-against) a constant. (Maybe you'd do this along the way...) <a href="http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/v/uniq_of_distns.pdf</a></p></li> <li><p>... this is not a stand-alone topic, but: the usual discussions of pseudo-differential operators (e.g., "symbols") somehow shrink from talking about <em>quotients</em> of TVS's in a grown-up way... If that hadn't been done earlier, and/or people had a (reasonable!) feeling of discomfort about the usual style of chatter in the psi-DO world, perhaps this could be happy-making.</p></li> <li><p>Meromorphic/holomorphic vector-valued functions (cf. Grothendieck c. 1953-4, and also various of my notes) e.g., meromorphic families of distributions... E.g., the $|x|^s$ family on $\mathbb R$ has <em>residues</em> which are the even-order derivatives of $\delta$, and ${\mathrm {sgn}}(x)\cdot |x|^s$ has as residues the odd-order derivatives of $\delta$. <a href="http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf" rel="noreferrer">http://www.math.umn.edu/~garrett/m/fun/Notes/09_vv_holo.pdf</a> Depending on context, there are somewhat-fancier things that I do find entertaining and also useful. Comments/correspondence are welcome.</p></li> </ul> Answer 17: <p>Two (<strong>edit:</strong> now four) not-so-usual examples come to my mind:</p> <ul> <li>There is the proof of the <em>central limit theorem</em> using Fourier analysis, as done in Chapter 7 of Hörmander's book. It's a cornerstone of probability theory and the Fourier-analytic proof presented by Hörmander is a nice showcase of the power of distribution theory: one gets a conceptually deep and useful theorem in full generality with relatively little effort;</li> <li>Something I have been meaning to try is to work out some "mathematical toy examples" of continuum limits of distributions supported on a lattice, as done e.g. in Chapter 15 of the book by R. Fernández, J. Fröhlich and A.D. Sokal, <em>Random Walks, Critical Phenomena, and Triviality in Quantum Field Theory</em> (Springer, 1992). I do not know if it is a feasible task, but since you are a specialist is constructive QFT, you should be able to do it... <strong>Edit 3:</strong> a classical and rather charming example of this has just came to my attention: the so-called <em>Bernstein polynomials</em>, used by their namesake Sergei N. Bernstein (the same mathematician who found the family of inequalities - also named after him - for smooth functions with compactly supported Fourier transform which underlie the Littlewood-Paley theory discussed in my Edit 1 below) to provide a probabilistic proof of the Weierstrass approximation theorem through the weak law of large numbers. Taking formal adjoints of the sequence of Bernstein polynomial operators provides a general lattice approximation for distributions on compact rectangular domains. A rather extensive discussion can be found in the short and beautiful book of the same title by G. G. Lorentz (the mathematician who invented the so-called Lorentz interpolation spaces, not the Dutch physicist).</li> <li><strong>(Edit 1)</strong> A third one, which I am particularly fond of, is the dyadic Fourier analysis of Littlewood and Paley, which is strongly connected to point (3). The connection to renormalization group ideas is quite obvious, and it has become a standard tool for obtaining important inequalities in nonlinear analysis (Gagliardo-Nirenberg, Moser, Schauder), all the way to the monumental Nash-Moser implicit function theorem (as shown by Hörmander - him again - in the 70's). A nice pedagogical exposition of these ideas can be found the book by S. Alinhac and P. Gérard, <em>Pseudo-Differential Operators and the Nash-Moser Theorem</em> (AMS, 2007) - just watch out for the numerous typographic errors!</li> <li><strong>(Edit 2)</strong> A fourth one, which is not actually unusual in itself but rather in its standpoint, is that many interesting examples of distributions (principal values, <em>partes finies</em>, extensions of homogeneous distributions to the origin, etc.) are obtained as extensions of distributions initially defined in (often finite-codimensional) subspaces of test functions, collectively called <em>renormalizations</em>. What would be interesting, also regarding applications to physics, is to treat such examples from this unified viewpoint. Indeed, as it can be seen e.g. by (Bogoliubov-Parasiuk-)Hepp's and Epstein-Glaser's treatments of perturbative renormalization in QFT, the latter really amounts to renormalizing distributions in the above sense. As a famous dictum of K. Hepp goes (<em>Proof of the Bogoliubov-Parasiuk Theorem on Renormalization</em>, Commun. Math. Phys. <strong>2</strong> (1966) 301-326): <blockquote> <p>"The renormalization theory (of Dyson) is in this framework a constructive form of the Hahn-Banach theorem." (pp. 302 - parentheses mine)</p> </blockquote></li> </ul> Answer 18: <p>One of my favourite applications of 'basic' distribution theory, which actually requires most of the tools in your list to be fully apreciated, is the Malgrange--Ehrenpreis Theorem on the local solvability of arbitrary constant coefficients PDE. There are a few proofs, but the proof in M.Taylor's book on Pseudodifferential Operators (last Section of first Chapter) is especially suited for your course I think. It is a substantial theorem though, so it might or might be not a good idea to include it, depending on the length and size of the course.</p> Answer 19: <p>No so much as applications, but I would love to learn about the basics of hyperfunctions (of the Japanese school) in such a course, if only in one dimension. Specifically I think it gives an easier way to define the wavefront set, which is what's required to explain why we can't square the delta function (referring to a comment above). See also the question <a href="https://mathoverflow.net/questions/48067/is-square-of-delta-function-defined-somewhere">Is square of Delta function defined somewhere?</a></p> <p>For what it's worth, I found the book</p> <ul> <li>Kato &amp; Struppa, <em>Fundamentals of Algebraic Microlocal Analysis</em></li> </ul> <p>particularly well written. The article in EoM is also great</p> <p><a href="https://www.encyclopediaofmath.org/index.php/Hyperfunction" rel="nofollow noreferrer">https://www.encyclopediaofmath.org/index.php/Hyperfunction</a></p> Answer 20: <p>Let me mention two nice applications to Sobolev spaces:</p> <p>The first result shows that if $f\in \mathcal{D}'$ and $\nabla f\in L^p_{\rm loc}$, then $f\in L^p_{\rm loc}$. For a short proof of this classical result see <a href="https://mathoverflow.net/a/296464/121665">https://mathoverflow.net/a/296464/121665</a></p> <p>For applications of this result to higher order Sobolev spaces see:<a href="https://mathoverflow.net/a/297392/121665">https://mathoverflow.net/a/297392/121665</a></p>
A roadmap for understanding perfectoid spaces
260,330
26
https://mathoverflow.net/questions/260330/a-roadmap-for-understanding-perfectoid-spaces
<p>Perfectoid spaces are this year's subject for the Arizona Winter School (<a href="http://swc.math.arizona.edu/" rel="noreferrer">link</a>) and, as preparation, I am currently trying to understand the subject better. There are wonderful explanatory accounts (<a href="http://www.ams.org/notices/201409/rnoti-p1082.pdf" rel="noreferrer">What is a perfectoid space</a>, <a href="https://mathoverflow.net/questions/65729/what-are-perfectoid-spaces">What are &quot;perfectoid spaces&quot;?</a>) but I would like to learn the technical details. An additional problem would be that I don't know much (anything) about rigid analytic geometry. I have found many lecture notes on the internet treating <a href="http://math.stanford.edu/~conrad/papers/aws.pdf" rel="noreferrer">non-archimedean geometry</a>, <a href="http://www-math.mit.edu/~poonen/STAGE-f13/Huber1993.pdf" rel="noreferrer">adic-spaces</a>, <a href="http://people.math.gatech.edu/~mbaker/pdf/aws07mb_v4.pdf" rel="noreferrer">Berkovich spaces</a> etc but I don't know where to start learning. Moreover, I don't know which things are important and which things can be skipped at first reading. I am familiar with schemes and local fields at the least. </p> <p>Is there a natural way to progress through the material?</p>
[ { "answer_id": 261712, "score": 11, "text": "<p>I can tell from personal experience that it is possible to learn perfectoid spaces without knowing rigid geometry, just like it is possible to learn schemes or even stacks without knowing much about varieties over complex numbers. In fact, it's even possible to successfully transition to research with this approach. Of course, for the approach to be meaningful/successful you need some \"mathematical maturity\" (in the sense of being able to clearly distinguish easy/formal parts of the theory from the real meat); for instance, you need a good command of commutative algebra/algebraic geometry (to say the least).</p>\n\n<p>If you want to pursue this, I suggest reading (more or less line by line) Wedhorn's \"Adic spaces\" for the basics and then (or in parallel) Scholze's \"Perfectoid spaces.\" </p>\n\n<p>Before wise elders start reprimanding me for giving such \"irresponsible\" advice, let me issue a couple of caveats:</p>\n\n<ol>\n<li>This approach is not for everyone. Specifically, if you feel you need \"motivating examples\" at every step and find it difficult to swallow dry theories for the mere sake of their own intrinsic beauty, then I think you're better off studying in a more linear fashion (i.e., learning rigid geometry first).</li>\n<li>You'll need to take some parts of section 2 of \"Perfectoid spaces\" on faith because you will not know much about Berkovich spaces (but the point is, they are not really needed!). For instance, you will need to ignore the description of the types of points on $(\\mathbb{P}^1)^{ad}$ (an overrated example anyway, from my somewhat limited experience).</li>\n<li>If you take this approach and want to at some later point transition to research, you have to be acutely aware of the fact that for a long time your intuition will derive from a good command of dry aspects of the theory (or from analogies with related theories that you know well) and not from examples or of concrete computations. In particular, you need to be aware of the danger of making big blunders when forming intuitions in this way about what should be true or how to prove something. Try to continuously bridge the gap as you move on.</li>\n</ol>\n\n<p>I do agree with the others though that it may be a little bit too late for the AWS to be entirely meaningful. Even if you do not completely understand the lectures, try to isolate the contact points with the material that you've been studying in order to get something out of them.</p>\n" }, { "answer_id": 261678, "score": 10, "text": "<p>Apart from all the written sources, there are six videos of Peter Scholze's lectures at the IHÉS in 2011 on Perfectoid Spaces and the Weight-Monodromy Conjecture : </p>\n\n<p><a href=\"http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech\">http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech</a></p>\n" }, { "answer_id": 261693, "score": 10, "text": "<p>I agree with nfdc23, notwithstanding here is some kind of 'roadmap' but I am by no means an expert.</p>\n\n<p>I would begin with Brian Conrad's \"Several approaches to Non-Archimedean geometry\" chapter in a set of lecture notes of a previous AWS <a href=\"http://www.ams.org/mathscinet-getitem?mr=MR2482345\">http://www.ams.org/mathscinet-getitem?mr=MR2482345</a> and there are many useful exercises there. Also see the second half of Tate's foreword and Berkovich's foreword to this book for quite an enlightening historical perspective.</p>\n\n<p>This doesn't treat adic geometry at all, but should give a good idea of why one might want a 'rigid analytic geometry' and the many many many subtleties that arise in the theories.</p>\n\n<p>For all the details in say a concrete case like the rigid projective line there is an e.g. first chapter in excellent textbook of Fresnel and van der Put <a href=\"http://www.ams.org/mathscinet-getitem?mr=MR2014891\">http://www.ams.org/mathscinet-getitem?mr=MR2014891</a> and the rest of the book has a lot of things that are difficult to find elsewhere. Note this isn't just a translation of the French version, there is a lot more in it than in the 1981 book.</p>\n\n<p>Then I suppose it is a matter of what you are interested in. For me it was things related to Picard--Fuchs, so the book of André was a treasure trove <a href=\"http://www.ams.org/mathscinet-getitem?mr=MR1978691\">http://www.ams.org/mathscinet-getitem?mr=MR1978691</a> using the Berkovich dialect.</p>\n\n<p>Then I guess for the adic case which isn't such a massive jump from Berkovich spaces, there are excellent lecture notes of Wedhorn (his webpage) and also its worth noting Huber was an extremely clear writer of articles, and for the perfectoid theory well you are lucky because Scholze is also a remarkably clear writer.</p>\n" }, { "answer_id": 261712, "score": 11, "text": "<p>I can tell from personal experience that it is possible to learn perfectoid spaces without knowing rigid geometry, just like it is possible to learn schemes or even stacks without knowing much about varieties over complex numbers. In fact, it's even possible to successfully transition to research with this approach. Of course, for the approach to be meaningful/successful you need some \"mathematical maturity\" (in the sense of being able to clearly distinguish easy/formal parts of the theory from the real meat); for instance, you need a good command of commutative algebra/algebraic geometry (to say the least).</p>\n\n<p>If you want to pursue this, I suggest reading (more or less line by line) Wedhorn's \"Adic spaces\" for the basics and then (or in parallel) Scholze's \"Perfectoid spaces.\" </p>\n\n<p>Before wise elders start reprimanding me for giving such \"irresponsible\" advice, let me issue a couple of caveats:</p>\n\n<ol>\n<li>This approach is not for everyone. Specifically, if you feel you need \"motivating examples\" at every step and find it difficult to swallow dry theories for the mere sake of their own intrinsic beauty, then I think you're better off studying in a more linear fashion (i.e., learning rigid geometry first).</li>\n<li>You'll need to take some parts of section 2 of \"Perfectoid spaces\" on faith because you will not know much about Berkovich spaces (but the point is, they are not really needed!). For instance, you will need to ignore the description of the types of points on $(\\mathbb{P}^1)^{ad}$ (an overrated example anyway, from my somewhat limited experience).</li>\n<li>If you take this approach and want to at some later point transition to research, you have to be acutely aware of the fact that for a long time your intuition will derive from a good command of dry aspects of the theory (or from analogies with related theories that you know well) and not from examples or of concrete computations. In particular, you need to be aware of the danger of making big blunders when forming intuitions in this way about what should be true or how to prove something. Try to continuously bridge the gap as you move on.</li>\n</ol>\n\n<p>I do agree with the others though that it may be a little bit too late for the AWS to be entirely meaningful. Even if you do not completely understand the lectures, try to isolate the contact points with the material that you've been studying in order to get something out of them.</p>\n" }, { "answer_id": 261678, "score": 10, "text": "<p>Apart from all the written sources, there are six videos of Peter Scholze's lectures at the IHÉS in 2011 on Perfectoid Spaces and the Weight-Monodromy Conjecture : </p>\n\n<p><a href=\"http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech\">http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech</a></p>\n" }, { "answer_id": 261693, "score": 10, "text": "<p>I agree with nfdc23, notwithstanding here is some kind of 'roadmap' but I am by no means an expert.</p>\n\n<p>I would begin with Brian Conrad's \"Several approaches to Non-Archimedean geometry\" chapter in a set of lecture notes of a previous AWS <a href=\"http://www.ams.org/mathscinet-getitem?mr=MR2482345\">http://www.ams.org/mathscinet-getitem?mr=MR2482345</a> and there are many useful exercises there. Also see the second half of Tate's foreword and Berkovich's foreword to this book for quite an enlightening historical perspective.</p>\n\n<p>This doesn't treat adic geometry at all, but should give a good idea of why one might want a 'rigid analytic geometry' and the many many many subtleties that arise in the theories.</p>\n\n<p>For all the details in say a concrete case like the rigid projective line there is an e.g. first chapter in excellent textbook of Fresnel and van der Put <a href=\"http://www.ams.org/mathscinet-getitem?mr=MR2014891\">http://www.ams.org/mathscinet-getitem?mr=MR2014891</a> and the rest of the book has a lot of things that are difficult to find elsewhere. Note this isn't just a translation of the French version, there is a lot more in it than in the 1981 book.</p>\n\n<p>Then I suppose it is a matter of what you are interested in. For me it was things related to Picard--Fuchs, so the book of André was a treasure trove <a href=\"http://www.ams.org/mathscinet-getitem?mr=MR1978691\">http://www.ams.org/mathscinet-getitem?mr=MR1978691</a> using the Berkovich dialect.</p>\n\n<p>Then I guess for the adic case which isn't such a massive jump from Berkovich spaces, there are excellent lecture notes of Wedhorn (his webpage) and also its worth noting Huber was an extremely clear writer of articles, and for the perfectoid theory well you are lucky because Scholze is also a remarkably clear writer.</p>\n" }, { "answer_id": 261712, "score": 11, "text": "<p>I can tell from personal experience that it is possible to learn perfectoid spaces without knowing rigid geometry, just like it is possible to learn schemes or even stacks without knowing much about varieties over complex numbers. In fact, it's even possible to successfully transition to research with this approach. Of course, for the approach to be meaningful/successful you need some \"mathematical maturity\" (in the sense of being able to clearly distinguish easy/formal parts of the theory from the real meat); for instance, you need a good command of commutative algebra/algebraic geometry (to say the least).</p>\n\n<p>If you want to pursue this, I suggest reading (more or less line by line) Wedhorn's \"Adic spaces\" for the basics and then (or in parallel) Scholze's \"Perfectoid spaces.\" </p>\n\n<p>Before wise elders start reprimanding me for giving such \"irresponsible\" advice, let me issue a couple of caveats:</p>\n\n<ol>\n<li>This approach is not for everyone. Specifically, if you feel you need \"motivating examples\" at every step and find it difficult to swallow dry theories for the mere sake of their own intrinsic beauty, then I think you're better off studying in a more linear fashion (i.e., learning rigid geometry first).</li>\n<li>You'll need to take some parts of section 2 of \"Perfectoid spaces\" on faith because you will not know much about Berkovich spaces (but the point is, they are not really needed!). For instance, you will need to ignore the description of the types of points on $(\\mathbb{P}^1)^{ad}$ (an overrated example anyway, from my somewhat limited experience).</li>\n<li>If you take this approach and want to at some later point transition to research, you have to be acutely aware of the fact that for a long time your intuition will derive from a good command of dry aspects of the theory (or from analogies with related theories that you know well) and not from examples or of concrete computations. In particular, you need to be aware of the danger of making big blunders when forming intuitions in this way about what should be true or how to prove something. Try to continuously bridge the gap as you move on.</li>\n</ol>\n\n<p>I do agree with the others though that it may be a little bit too late for the AWS to be entirely meaningful. Even if you do not completely understand the lectures, try to isolate the contact points with the material that you've been studying in order to get something out of them.</p>\n" }, { "answer_id": 261678, "score": 10, "text": "<p>Apart from all the written sources, there are six videos of Peter Scholze's lectures at the IHÉS in 2011 on Perfectoid Spaces and the Weight-Monodromy Conjecture : </p>\n\n<p><a href=\"http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech\">http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech</a></p>\n" }, { "answer_id": 261693, "score": 10, "text": "<p>I agree with nfdc23, notwithstanding here is some kind of 'roadmap' but I am by no means an expert.</p>\n\n<p>I would begin with Brian Conrad's \"Several approaches to Non-Archimedean geometry\" chapter in a set of lecture notes of a previous AWS <a href=\"http://www.ams.org/mathscinet-getitem?mr=MR2482345\">http://www.ams.org/mathscinet-getitem?mr=MR2482345</a> and there are many useful exercises there. Also see the second half of Tate's foreword and Berkovich's foreword to this book for quite an enlightening historical perspective.</p>\n\n<p>This doesn't treat adic geometry at all, but should give a good idea of why one might want a 'rigid analytic geometry' and the many many many subtleties that arise in the theories.</p>\n\n<p>For all the details in say a concrete case like the rigid projective line there is an e.g. first chapter in excellent textbook of Fresnel and van der Put <a href=\"http://www.ams.org/mathscinet-getitem?mr=MR2014891\">http://www.ams.org/mathscinet-getitem?mr=MR2014891</a> and the rest of the book has a lot of things that are difficult to find elsewhere. Note this isn't just a translation of the French version, there is a lot more in it than in the 1981 book.</p>\n\n<p>Then I suppose it is a matter of what you are interested in. For me it was things related to Picard--Fuchs, so the book of André was a treasure trove <a href=\"http://www.ams.org/mathscinet-getitem?mr=MR1978691\">http://www.ams.org/mathscinet-getitem?mr=MR1978691</a> using the Berkovich dialect.</p>\n\n<p>Then I guess for the adic case which isn't such a massive jump from Berkovich spaces, there are excellent lecture notes of Wedhorn (his webpage) and also its worth noting Huber was an extremely clear writer of articles, and for the perfectoid theory well you are lucky because Scholze is also a remarkably clear writer.</p>\n" }, { "answer_id": 261712, "score": 11, "text": "<p>I can tell from personal experience that it is possible to learn perfectoid spaces without knowing rigid geometry, just like it is possible to learn schemes or even stacks without knowing much about varieties over complex numbers. In fact, it's even possible to successfully transition to research with this approach. Of course, for the approach to be meaningful/successful you need some \"mathematical maturity\" (in the sense of being able to clearly distinguish easy/formal parts of the theory from the real meat); for instance, you need a good command of commutative algebra/algebraic geometry (to say the least).</p>\n\n<p>If you want to pursue this, I suggest reading (more or less line by line) Wedhorn's \"Adic spaces\" for the basics and then (or in parallel) Scholze's \"Perfectoid spaces.\" </p>\n\n<p>Before wise elders start reprimanding me for giving such \"irresponsible\" advice, let me issue a couple of caveats:</p>\n\n<ol>\n<li>This approach is not for everyone. Specifically, if you feel you need \"motivating examples\" at every step and find it difficult to swallow dry theories for the mere sake of their own intrinsic beauty, then I think you're better off studying in a more linear fashion (i.e., learning rigid geometry first).</li>\n<li>You'll need to take some parts of section 2 of \"Perfectoid spaces\" on faith because you will not know much about Berkovich spaces (but the point is, they are not really needed!). For instance, you will need to ignore the description of the types of points on $(\\mathbb{P}^1)^{ad}$ (an overrated example anyway, from my somewhat limited experience).</li>\n<li>If you take this approach and want to at some later point transition to research, you have to be acutely aware of the fact that for a long time your intuition will derive from a good command of dry aspects of the theory (or from analogies with related theories that you know well) and not from examples or of concrete computations. In particular, you need to be aware of the danger of making big blunders when forming intuitions in this way about what should be true or how to prove something. Try to continuously bridge the gap as you move on.</li>\n</ol>\n\n<p>I do agree with the others though that it may be a little bit too late for the AWS to be entirely meaningful. Even if you do not completely understand the lectures, try to isolate the contact points with the material that you've been studying in order to get something out of them.</p>\n" }, { "answer_id": 261678, "score": 10, "text": "<p>Apart from all the written sources, there are six videos of Peter Scholze's lectures at the IHÉS in 2011 on Perfectoid Spaces and the Weight-Monodromy Conjecture : </p>\n\n<p><a href=\"http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech\">http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech</a></p>\n\n<p><a href=\"http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech\">http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech</a></p>\n" }, { "answer_id": 261693, "score": 10, "text": "<p>I agree with nfdc23, notwithstanding here is some kind of 'roadmap' but I am by no means an expert.</p>\n\n<p>I would begin with Brian Conrad's \"Several approaches to Non-Archimedean geometry\" chapter in a set of lecture notes of a previous AWS <a href=\"http://www.ams.org/mathscinet-getitem?mr=MR2482345\">http://www.ams.org/mathscinet-getitem?mr=MR2482345</a> and there are many useful exercises there. Also see the second half of Tate's foreword and Berkovich's foreword to this book for quite an enlightening historical perspective.</p>\n\n<p>This doesn't treat adic geometry at all, but should give a good idea of why one might want a 'rigid analytic geometry' and the many many many subtleties that arise in the theories.</p>\n\n<p>For all the details in say a concrete case like the rigid projective line there is an e.g. first chapter in excellent textbook of Fresnel and van der Put <a href=\"http://www.ams.org/mathscinet-getitem?mr=MR2014891\">http://www.ams.org/mathscinet-getitem?mr=MR2014891</a> and the rest of the book has a lot of things that are difficult to find elsewhere. Note this isn't just a translation of the French version, there is a lot more in it than in the 1981 book.</p>\n\n<p>Then I suppose it is a matter of what you are interested in. For me it was things related to Picard--Fuchs, so the book of André was a treasure trove <a href=\"http://www.ams.org/mathscinet-getitem?mr=MR1978691\">http://www.ams.org/mathscinet-getitem?mr=MR1978691</a> using the Berkovich dialect.</p>\n\n<p>Then I guess for the adic case which isn't such a massive jump from Berkovich spaces, there are excellent lecture notes of Wedhorn (his webpage) and also its worth noting Huber was an extremely clear writer of articles, and for the perfectoid theory well you are lucky because Scholze is also a remarkably clear writer.</p>\n" } ]
Question: A roadmap for understanding perfectoid spaces Answer 1: <p>I can tell from personal experience that it is possible to learn perfectoid spaces without knowing rigid geometry, just like it is possible to learn schemes or even stacks without knowing much about varieties over complex numbers. In fact, it's even possible to successfully transition to research with this approach. Of course, for the approach to be meaningful/successful you need some "mathematical maturity" (in the sense of being able to clearly distinguish easy/formal parts of the theory from the real meat); for instance, you need a good command of commutative algebra/algebraic geometry (to say the least).</p> <p>If you want to pursue this, I suggest reading (more or less line by line) Wedhorn's "Adic spaces" for the basics and then (or in parallel) Scholze's "Perfectoid spaces." </p> <p>Before wise elders start reprimanding me for giving such "irresponsible" advice, let me issue a couple of caveats:</p> <ol> <li>This approach is not for everyone. Specifically, if you feel you need "motivating examples" at every step and find it difficult to swallow dry theories for the mere sake of their own intrinsic beauty, then I think you're better off studying in a more linear fashion (i.e., learning rigid geometry first).</li> <li>You'll need to take some parts of section 2 of "Perfectoid spaces" on faith because you will not know much about Berkovich spaces (but the point is, they are not really needed!). For instance, you will need to ignore the description of the types of points on $(\mathbb{P}^1)^{ad}$ (an overrated example anyway, from my somewhat limited experience).</li> <li>If you take this approach and want to at some later point transition to research, you have to be acutely aware of the fact that for a long time your intuition will derive from a good command of dry aspects of the theory (or from analogies with related theories that you know well) and not from examples or of concrete computations. In particular, you need to be aware of the danger of making big blunders when forming intuitions in this way about what should be true or how to prove something. Try to continuously bridge the gap as you move on.</li> </ol> <p>I do agree with the others though that it may be a little bit too late for the AWS to be entirely meaningful. Even if you do not completely understand the lectures, try to isolate the contact points with the material that you've been studying in order to get something out of them.</p> Answer 2: <p>Apart from all the written sources, there are six videos of Peter Scholze's lectures at the IHÉS in 2011 on Perfectoid Spaces and the Weight-Monodromy Conjecture : </p> <p><a href="http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech">http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech">http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech">http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech">http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech">http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech">http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech</a></p> Answer 3: <p>I agree with nfdc23, notwithstanding here is some kind of 'roadmap' but I am by no means an expert.</p> <p>I would begin with Brian Conrad's "Several approaches to Non-Archimedean geometry" chapter in a set of lecture notes of a previous AWS <a href="http://www.ams.org/mathscinet-getitem?mr=MR2482345">http://www.ams.org/mathscinet-getitem?mr=MR2482345</a> and there are many useful exercises there. Also see the second half of Tate's foreword and Berkovich's foreword to this book for quite an enlightening historical perspective.</p> <p>This doesn't treat adic geometry at all, but should give a good idea of why one might want a 'rigid analytic geometry' and the many many many subtleties that arise in the theories.</p> <p>For all the details in say a concrete case like the rigid projective line there is an e.g. first chapter in excellent textbook of Fresnel and van der Put <a href="http://www.ams.org/mathscinet-getitem?mr=MR2014891">http://www.ams.org/mathscinet-getitem?mr=MR2014891</a> and the rest of the book has a lot of things that are difficult to find elsewhere. Note this isn't just a translation of the French version, there is a lot more in it than in the 1981 book.</p> <p>Then I suppose it is a matter of what you are interested in. For me it was things related to Picard--Fuchs, so the book of André was a treasure trove <a href="http://www.ams.org/mathscinet-getitem?mr=MR1978691">http://www.ams.org/mathscinet-getitem?mr=MR1978691</a> using the Berkovich dialect.</p> <p>Then I guess for the adic case which isn't such a massive jump from Berkovich spaces, there are excellent lecture notes of Wedhorn (his webpage) and also its worth noting Huber was an extremely clear writer of articles, and for the perfectoid theory well you are lucky because Scholze is also a remarkably clear writer.</p> Answer 4: <p>I can tell from personal experience that it is possible to learn perfectoid spaces without knowing rigid geometry, just like it is possible to learn schemes or even stacks without knowing much about varieties over complex numbers. In fact, it's even possible to successfully transition to research with this approach. Of course, for the approach to be meaningful/successful you need some "mathematical maturity" (in the sense of being able to clearly distinguish easy/formal parts of the theory from the real meat); for instance, you need a good command of commutative algebra/algebraic geometry (to say the least).</p> <p>If you want to pursue this, I suggest reading (more or less line by line) Wedhorn's "Adic spaces" for the basics and then (or in parallel) Scholze's "Perfectoid spaces." </p> <p>Before wise elders start reprimanding me for giving such "irresponsible" advice, let me issue a couple of caveats:</p> <ol> <li>This approach is not for everyone. Specifically, if you feel you need "motivating examples" at every step and find it difficult to swallow dry theories for the mere sake of their own intrinsic beauty, then I think you're better off studying in a more linear fashion (i.e., learning rigid geometry first).</li> <li>You'll need to take some parts of section 2 of "Perfectoid spaces" on faith because you will not know much about Berkovich spaces (but the point is, they are not really needed!). For instance, you will need to ignore the description of the types of points on $(\mathbb{P}^1)^{ad}$ (an overrated example anyway, from my somewhat limited experience).</li> <li>If you take this approach and want to at some later point transition to research, you have to be acutely aware of the fact that for a long time your intuition will derive from a good command of dry aspects of the theory (or from analogies with related theories that you know well) and not from examples or of concrete computations. In particular, you need to be aware of the danger of making big blunders when forming intuitions in this way about what should be true or how to prove something. Try to continuously bridge the gap as you move on.</li> </ol> <p>I do agree with the others though that it may be a little bit too late for the AWS to be entirely meaningful. Even if you do not completely understand the lectures, try to isolate the contact points with the material that you've been studying in order to get something out of them.</p> Answer 5: <p>Apart from all the written sources, there are six videos of Peter Scholze's lectures at the IHÉS in 2011 on Perfectoid Spaces and the Weight-Monodromy Conjecture : </p> <p><a href="http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech">http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech">http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech">http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech">http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech">http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech">http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech</a></p> Answer 6: <p>I agree with nfdc23, notwithstanding here is some kind of 'roadmap' but I am by no means an expert.</p> <p>I would begin with Brian Conrad's "Several approaches to Non-Archimedean geometry" chapter in a set of lecture notes of a previous AWS <a href="http://www.ams.org/mathscinet-getitem?mr=MR2482345">http://www.ams.org/mathscinet-getitem?mr=MR2482345</a> and there are many useful exercises there. Also see the second half of Tate's foreword and Berkovich's foreword to this book for quite an enlightening historical perspective.</p> <p>This doesn't treat adic geometry at all, but should give a good idea of why one might want a 'rigid analytic geometry' and the many many many subtleties that arise in the theories.</p> <p>For all the details in say a concrete case like the rigid projective line there is an e.g. first chapter in excellent textbook of Fresnel and van der Put <a href="http://www.ams.org/mathscinet-getitem?mr=MR2014891">http://www.ams.org/mathscinet-getitem?mr=MR2014891</a> and the rest of the book has a lot of things that are difficult to find elsewhere. Note this isn't just a translation of the French version, there is a lot more in it than in the 1981 book.</p> <p>Then I suppose it is a matter of what you are interested in. For me it was things related to Picard--Fuchs, so the book of André was a treasure trove <a href="http://www.ams.org/mathscinet-getitem?mr=MR1978691">http://www.ams.org/mathscinet-getitem?mr=MR1978691</a> using the Berkovich dialect.</p> <p>Then I guess for the adic case which isn't such a massive jump from Berkovich spaces, there are excellent lecture notes of Wedhorn (his webpage) and also its worth noting Huber was an extremely clear writer of articles, and for the perfectoid theory well you are lucky because Scholze is also a remarkably clear writer.</p> Answer 7: <p>I can tell from personal experience that it is possible to learn perfectoid spaces without knowing rigid geometry, just like it is possible to learn schemes or even stacks without knowing much about varieties over complex numbers. In fact, it's even possible to successfully transition to research with this approach. Of course, for the approach to be meaningful/successful you need some "mathematical maturity" (in the sense of being able to clearly distinguish easy/formal parts of the theory from the real meat); for instance, you need a good command of commutative algebra/algebraic geometry (to say the least).</p> <p>If you want to pursue this, I suggest reading (more or less line by line) Wedhorn's "Adic spaces" for the basics and then (or in parallel) Scholze's "Perfectoid spaces." </p> <p>Before wise elders start reprimanding me for giving such "irresponsible" advice, let me issue a couple of caveats:</p> <ol> <li>This approach is not for everyone. Specifically, if you feel you need "motivating examples" at every step and find it difficult to swallow dry theories for the mere sake of their own intrinsic beauty, then I think you're better off studying in a more linear fashion (i.e., learning rigid geometry first).</li> <li>You'll need to take some parts of section 2 of "Perfectoid spaces" on faith because you will not know much about Berkovich spaces (but the point is, they are not really needed!). For instance, you will need to ignore the description of the types of points on $(\mathbb{P}^1)^{ad}$ (an overrated example anyway, from my somewhat limited experience).</li> <li>If you take this approach and want to at some later point transition to research, you have to be acutely aware of the fact that for a long time your intuition will derive from a good command of dry aspects of the theory (or from analogies with related theories that you know well) and not from examples or of concrete computations. In particular, you need to be aware of the danger of making big blunders when forming intuitions in this way about what should be true or how to prove something. Try to continuously bridge the gap as you move on.</li> </ol> <p>I do agree with the others though that it may be a little bit too late for the AWS to be entirely meaningful. Even if you do not completely understand the lectures, try to isolate the contact points with the material that you've been studying in order to get something out of them.</p> Answer 8: <p>Apart from all the written sources, there are six videos of Peter Scholze's lectures at the IHÉS in 2011 on Perfectoid Spaces and the Weight-Monodromy Conjecture : </p> <p><a href="http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech">http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech">http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech">http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech">http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech">http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech">http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech</a></p> Answer 9: <p>I agree with nfdc23, notwithstanding here is some kind of 'roadmap' but I am by no means an expert.</p> <p>I would begin with Brian Conrad's "Several approaches to Non-Archimedean geometry" chapter in a set of lecture notes of a previous AWS <a href="http://www.ams.org/mathscinet-getitem?mr=MR2482345">http://www.ams.org/mathscinet-getitem?mr=MR2482345</a> and there are many useful exercises there. Also see the second half of Tate's foreword and Berkovich's foreword to this book for quite an enlightening historical perspective.</p> <p>This doesn't treat adic geometry at all, but should give a good idea of why one might want a 'rigid analytic geometry' and the many many many subtleties that arise in the theories.</p> <p>For all the details in say a concrete case like the rigid projective line there is an e.g. first chapter in excellent textbook of Fresnel and van der Put <a href="http://www.ams.org/mathscinet-getitem?mr=MR2014891">http://www.ams.org/mathscinet-getitem?mr=MR2014891</a> and the rest of the book has a lot of things that are difficult to find elsewhere. Note this isn't just a translation of the French version, there is a lot more in it than in the 1981 book.</p> <p>Then I suppose it is a matter of what you are interested in. For me it was things related to Picard--Fuchs, so the book of André was a treasure trove <a href="http://www.ams.org/mathscinet-getitem?mr=MR1978691">http://www.ams.org/mathscinet-getitem?mr=MR1978691</a> using the Berkovich dialect.</p> <p>Then I guess for the adic case which isn't such a massive jump from Berkovich spaces, there are excellent lecture notes of Wedhorn (his webpage) and also its worth noting Huber was an extremely clear writer of articles, and for the perfectoid theory well you are lucky because Scholze is also a remarkably clear writer.</p> Answer 10: <p>I can tell from personal experience that it is possible to learn perfectoid spaces without knowing rigid geometry, just like it is possible to learn schemes or even stacks without knowing much about varieties over complex numbers. In fact, it's even possible to successfully transition to research with this approach. Of course, for the approach to be meaningful/successful you need some "mathematical maturity" (in the sense of being able to clearly distinguish easy/formal parts of the theory from the real meat); for instance, you need a good command of commutative algebra/algebraic geometry (to say the least).</p> <p>If you want to pursue this, I suggest reading (more or less line by line) Wedhorn's "Adic spaces" for the basics and then (or in parallel) Scholze's "Perfectoid spaces." </p> <p>Before wise elders start reprimanding me for giving such "irresponsible" advice, let me issue a couple of caveats:</p> <ol> <li>This approach is not for everyone. Specifically, if you feel you need "motivating examples" at every step and find it difficult to swallow dry theories for the mere sake of their own intrinsic beauty, then I think you're better off studying in a more linear fashion (i.e., learning rigid geometry first).</li> <li>You'll need to take some parts of section 2 of "Perfectoid spaces" on faith because you will not know much about Berkovich spaces (but the point is, they are not really needed!). For instance, you will need to ignore the description of the types of points on $(\mathbb{P}^1)^{ad}$ (an overrated example anyway, from my somewhat limited experience).</li> <li>If you take this approach and want to at some later point transition to research, you have to be acutely aware of the fact that for a long time your intuition will derive from a good command of dry aspects of the theory (or from analogies with related theories that you know well) and not from examples or of concrete computations. In particular, you need to be aware of the danger of making big blunders when forming intuitions in this way about what should be true or how to prove something. Try to continuously bridge the gap as you move on.</li> </ol> <p>I do agree with the others though that it may be a little bit too late for the AWS to be entirely meaningful. Even if you do not completely understand the lectures, try to isolate the contact points with the material that you've been studying in order to get something out of them.</p> Answer 11: <p>Apart from all the written sources, there are six videos of Peter Scholze's lectures at the IHÉS in 2011 on Perfectoid Spaces and the Weight-Monodromy Conjecture : </p> <p><a href="http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech">http://www.dailymotion.com/video/xm4b56_perfectoid-spaces-and-the-weight-monodromy-conjecture-1-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech">http://www.dailymotion.com/video/xmes49_perfectoid-spaces-and-the-weight-monodromy-conjecture-2-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech">http://www.dailymotion.com/video/xmet75_perfectoid-spaces-and-the-weight-monodromy-conjecture-3-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech">http://www.dailymotion.com/video/xmev2o_perfectoid-spaces-and-the-weight-monodromy-conjecture-4-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech">http://www.dailymotion.com/video/xmewlx_perfectoid-spaces-and-the-weight-monodromy-conjecture-5-6-peter-scholze_tech</a></p> <p><a href="http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech">http://www.dailymotion.com/video/xmip3g_perfectoid-spaces-and-the-weight-monodromy-conjecture-6-6-peter-scholze_tech</a></p> Answer 12: <p>I agree with nfdc23, notwithstanding here is some kind of 'roadmap' but I am by no means an expert.</p> <p>I would begin with Brian Conrad's "Several approaches to Non-Archimedean geometry" chapter in a set of lecture notes of a previous AWS <a href="http://www.ams.org/mathscinet-getitem?mr=MR2482345">http://www.ams.org/mathscinet-getitem?mr=MR2482345</a> and there are many useful exercises there. Also see the second half of Tate's foreword and Berkovich's foreword to this book for quite an enlightening historical perspective.</p> <p>This doesn't treat adic geometry at all, but should give a good idea of why one might want a 'rigid analytic geometry' and the many many many subtleties that arise in the theories.</p> <p>For all the details in say a concrete case like the rigid projective line there is an e.g. first chapter in excellent textbook of Fresnel and van der Put <a href="http://www.ams.org/mathscinet-getitem?mr=MR2014891">http://www.ams.org/mathscinet-getitem?mr=MR2014891</a> and the rest of the book has a lot of things that are difficult to find elsewhere. Note this isn't just a translation of the French version, there is a lot more in it than in the 1981 book.</p> <p>Then I suppose it is a matter of what you are interested in. For me it was things related to Picard--Fuchs, so the book of André was a treasure trove <a href="http://www.ams.org/mathscinet-getitem?mr=MR1978691">http://www.ams.org/mathscinet-getitem?mr=MR1978691</a> using the Berkovich dialect.</p> <p>Then I guess for the adic case which isn't such a massive jump from Berkovich spaces, there are excellent lecture notes of Wedhorn (his webpage) and also its worth noting Huber was an extremely clear writer of articles, and for the perfectoid theory well you are lucky because Scholze is also a remarkably clear writer.</p>
Research directions in persistent homology
260,060
24
https://mathoverflow.net/questions/260060/research-directions-in-persistent-homology
<p>I am interested in what are the possible directions for new research in persistent homology (more of the mathematical theoretical aspects rather than the computer algorithm aspects).</p> <p>So far from googling, proving stability (not affected by small changes) seems to be one direction.</p> <p>Another direction I have seen is the inverse problem of persistent homology: to what extent can the original space be reconstructed from the persistent homology.</p> <blockquote> <p>Are there any other research or generalizations of persistent homology?</p> </blockquote>
[ { "answer_id": 260062, "score": 19, "text": "<p>Persistent homology has applications in symplectic geometry. For one striking example, see the recent paper <a href=\"https://arxiv.org/abs/1412.8277\">Autonomous Hamiltonian flows, Hofer's geometry and persistence modules</a> by Polterovich and Shelukhin. They use persistent homology to give very explicit examples of area-preserving diffeomorphisms of closed surfaces which cannot be included into $1$-parameter subgroups of such.</p>\n" }, { "answer_id": 260078, "score": 15, "text": "<p>Persistent homology is related to spectral sequences. There is already a discussion <a href=\"https://mathoverflow.net/questions/208406/persistence-barcodes-and-spectral-sequences\">Persistence barcodes and spectral sequences</a> about if this leads to new theoretical insides and there are stated further references (for example, there is stated the book of Edelsbrunner and Harer, which is a standard book for Persistent homology and can find further discussions about some research directions in Edelsbrunner and Harer's works as well).</p>\n\n<p>Furthermore, Persistent homology leads to discrete Morse theory, what can be used for computing cellular sheaf cohomology (if you want to know more about it, i.e. Justin Curry's work could be interesting for you). </p>\n" }, { "answer_id": 260062, "score": 19, "text": "<p>Persistent homology has applications in symplectic geometry. For one striking example, see the recent paper <a href=\"https://arxiv.org/abs/1412.8277\">Autonomous Hamiltonian flows, Hofer's geometry and persistence modules</a> by Polterovich and Shelukhin. They use persistent homology to give very explicit examples of area-preserving diffeomorphisms of closed surfaces which cannot be included into $1$-parameter subgroups of such.</p>\n" }, { "answer_id": 260078, "score": 15, "text": "<p>Persistent homology is related to spectral sequences. There is already a discussion <a href=\"https://mathoverflow.net/questions/208406/persistence-barcodes-and-spectral-sequences\">Persistence barcodes and spectral sequences</a> about if this leads to new theoretical insides and there are stated further references (for example, there is stated the book of Edelsbrunner and Harer, which is a standard book for Persistent homology and can find further discussions about some research directions in Edelsbrunner and Harer's works as well).</p>\n\n<p>Furthermore, Persistent homology leads to discrete Morse theory, what can be used for computing cellular sheaf cohomology (if you want to know more about it, i.e. Justin Curry's work could be interesting for you). </p>\n" }, { "answer_id": 260062, "score": 19, "text": "<p>Persistent homology has applications in symplectic geometry. For one striking example, see the recent paper <a href=\"https://arxiv.org/abs/1412.8277\">Autonomous Hamiltonian flows, Hofer's geometry and persistence modules</a> by Polterovich and Shelukhin. They use persistent homology to give very explicit examples of area-preserving diffeomorphisms of closed surfaces which cannot be included into $1$-parameter subgroups of such.</p>\n" }, { "answer_id": 260078, "score": 15, "text": "<p>Persistent homology is related to spectral sequences. There is already a discussion <a href=\"https://mathoverflow.net/questions/208406/persistence-barcodes-and-spectral-sequences\">Persistence barcodes and spectral sequences</a> about if this leads to new theoretical insides and there are stated further references (for example, there is stated the book of Edelsbrunner and Harer, which is a standard book for Persistent homology and can find further discussions about some research directions in Edelsbrunner and Harer's works as well).</p>\n\n<p>Furthermore, Persistent homology leads to discrete Morse theory, what can be used for computing cellular sheaf cohomology (if you want to know more about it, i.e. Justin Curry's work could be interesting for you). </p>\n" }, { "answer_id": 260062, "score": 19, "text": "<p>Persistent homology has applications in symplectic geometry. For one striking example, see the recent paper <a href=\"https://arxiv.org/abs/1412.8277\">Autonomous Hamiltonian flows, Hofer's geometry and persistence modules</a> by Polterovich and Shelukhin. They use persistent homology to give very explicit examples of area-preserving diffeomorphisms of closed surfaces which cannot be included into $1$-parameter subgroups of such.</p>\n" }, { "answer_id": 260078, "score": 15, "text": "<p>Persistent homology is related to spectral sequences. There is already a discussion <a href=\"https://mathoverflow.net/questions/208406/persistence-barcodes-and-spectral-sequences\">Persistence barcodes and spectral sequences</a> about if this leads to new theoretical insides and there are stated further references (for example, there is stated the book of Edelsbrunner and Harer, which is a standard book for Persistent homology and can find further discussions about some research directions in Edelsbrunner and Harer's works as well).</p>\n\n<p>Furthermore, Persistent homology leads to discrete Morse theory, what can be used for computing cellular sheaf cohomology (if you want to know more about it, i.e. Justin Curry's work could be interesting for you). </p>\n" } ]
Question: Research directions in persistent homology Answer 1: <p>Persistent homology has applications in symplectic geometry. For one striking example, see the recent paper <a href="https://arxiv.org/abs/1412.8277">Autonomous Hamiltonian flows, Hofer's geometry and persistence modules</a> by Polterovich and Shelukhin. They use persistent homology to give very explicit examples of area-preserving diffeomorphisms of closed surfaces which cannot be included into $1$-parameter subgroups of such.</p> Answer 2: <p>Persistent homology is related to spectral sequences. There is already a discussion <a href="https://mathoverflow.net/questions/208406/persistence-barcodes-and-spectral-sequences">Persistence barcodes and spectral sequences</a> about if this leads to new theoretical insides and there are stated further references (for example, there is stated the book of Edelsbrunner and Harer, which is a standard book for Persistent homology and can find further discussions about some research directions in Edelsbrunner and Harer's works as well).</p> <p>Furthermore, Persistent homology leads to discrete Morse theory, what can be used for computing cellular sheaf cohomology (if you want to know more about it, i.e. Justin Curry's work could be interesting for you). </p> Answer 3: <p>Persistent homology has applications in symplectic geometry. For one striking example, see the recent paper <a href="https://arxiv.org/abs/1412.8277">Autonomous Hamiltonian flows, Hofer's geometry and persistence modules</a> by Polterovich and Shelukhin. They use persistent homology to give very explicit examples of area-preserving diffeomorphisms of closed surfaces which cannot be included into $1$-parameter subgroups of such.</p> Answer 4: <p>Persistent homology is related to spectral sequences. There is already a discussion <a href="https://mathoverflow.net/questions/208406/persistence-barcodes-and-spectral-sequences">Persistence barcodes and spectral sequences</a> about if this leads to new theoretical insides and there are stated further references (for example, there is stated the book of Edelsbrunner and Harer, which is a standard book for Persistent homology and can find further discussions about some research directions in Edelsbrunner and Harer's works as well).</p> <p>Furthermore, Persistent homology leads to discrete Morse theory, what can be used for computing cellular sheaf cohomology (if you want to know more about it, i.e. Justin Curry's work could be interesting for you). </p> Answer 5: <p>Persistent homology has applications in symplectic geometry. For one striking example, see the recent paper <a href="https://arxiv.org/abs/1412.8277">Autonomous Hamiltonian flows, Hofer's geometry and persistence modules</a> by Polterovich and Shelukhin. They use persistent homology to give very explicit examples of area-preserving diffeomorphisms of closed surfaces which cannot be included into $1$-parameter subgroups of such.</p> Answer 6: <p>Persistent homology is related to spectral sequences. There is already a discussion <a href="https://mathoverflow.net/questions/208406/persistence-barcodes-and-spectral-sequences">Persistence barcodes and spectral sequences</a> about if this leads to new theoretical insides and there are stated further references (for example, there is stated the book of Edelsbrunner and Harer, which is a standard book for Persistent homology and can find further discussions about some research directions in Edelsbrunner and Harer's works as well).</p> <p>Furthermore, Persistent homology leads to discrete Morse theory, what can be used for computing cellular sheaf cohomology (if you want to know more about it, i.e. Justin Curry's work could be interesting for you). </p> Answer 7: <p>Persistent homology has applications in symplectic geometry. For one striking example, see the recent paper <a href="https://arxiv.org/abs/1412.8277">Autonomous Hamiltonian flows, Hofer's geometry and persistence modules</a> by Polterovich and Shelukhin. They use persistent homology to give very explicit examples of area-preserving diffeomorphisms of closed surfaces which cannot be included into $1$-parameter subgroups of such.</p> Answer 8: <p>Persistent homology is related to spectral sequences. There is already a discussion <a href="https://mathoverflow.net/questions/208406/persistence-barcodes-and-spectral-sequences">Persistence barcodes and spectral sequences</a> about if this leads to new theoretical insides and there are stated further references (for example, there is stated the book of Edelsbrunner and Harer, which is a standard book for Persistent homology and can find further discussions about some research directions in Edelsbrunner and Harer's works as well).</p> <p>Furthermore, Persistent homology leads to discrete Morse theory, what can be used for computing cellular sheaf cohomology (if you want to know more about it, i.e. Justin Curry's work could be interesting for you). </p>
What is the set of all &quot;pseudo-rational&quot; numbers (see details)?
259,565
21
https://mathoverflow.net/questions/259565/what-is-the-set-of-all-pseudo-rational-numbers-see-details
<p>Define a “pseudo-rational” number to be a real number $q$ that can be written as</p> <p>$q=\sum_{n=1}^{\infty} \frac{P(n)}{Q(n)}$</p> <p>Where $P(x)$ and $Q(x)$ are fixed integer polynomials (independent of n). All rational numbers are pseudo-rational, as is $\pi^2$ using $P(x)=6,Q(x)=x^2$. There must exist numbers that are not pseudo-rational (defined as "pseudo-irrational") because the set of pseudo-rationals is countable. Is $e$ pseudo-rational? Is $\sqrt{2}$?</p>
[ { "answer_id": 259708, "score": 11, "text": "<p><em>One partial answer:</em></p>\n\n<p>$\\pi$ and $\\ln(2)$ are both pseudo-rational:\n\\begin{align}\n\\ln(2) &amp;= \\sum\\frac{1}{2n\\,(2n-1)} \\\\\n\\pi &amp;= \\sum\\frac{3}{n\\,(2n-1)\\,(4n-3)} \\\\\n\\end{align}</p>\n\n<p>These follow from statements in <a href=\"https://en.wikipedia.org/wiki/Digamma_function\">Wikipedia</a>, including Gauss's digamma theorem, and are also asserted by Mathematica. By similar manipulations, $\\pi\\sqrt{3}$ and $\\ln(3)$ are pseudo-rational also.</p>\n\n<p>$\\ $</p>\n\n<p><em>Some questions collected from the comments:</em></p>\n\n<ul>\n<li>Are the pseudo-rationals closed under multiplication?</li>\n<li>Are all pseudo-rationals <a href=\"https://en.wikipedia.org/wiki/Ring_of_periods\">periods</a>?</li>\n<li>Is there a procedure to decide if the pseudo-rational from $P,Q$ is positive? </li>\n</ul>\n\n<p>The last one is difficult because there are non-trivial zeroes like $\\sum\\left(\\dfrac{6}{n^2}-\\dfrac{8}{(2n-1)^2}\\right)$.</p>\n" }, { "answer_id": 259780, "score": 7, "text": "<p>Also possibly of interest:</p>\n\n<p>$$ \\sum_{n=1}^\\infty \\left(-\\frac{2}{b(n-1/b)(n+1/b)} + \\frac{b}{n(n+1)}\\right) = \\pi \\cot(\\pi/b)$$</p>\n\n<p>$$ \\sum_{n=1}^\\infty {\\frac {t \\left( {t}^{2}{n}^{2}+2\\,{n}^{2}+2\\,n+1 \\right) }{ \\left( n+\n1 \\right) n \\left( {t}^{2}{n}^{2}+1 \\right) }}\n = \\pi \\coth(\\pi/t) $$</p>\n\n<p>EDIT: \nAnd, if I'm not mistaken,</p>\n\n<p>$$ \\sum_{n=1}^\\infty \\left(\\frac{1-m}{mn} + \\sum_{k=1}^{m-1} \\frac{1}{mn-k}\\right) = \\ln(m) $$\nfor positive integers $m$.</p>\n" }, { "answer_id": 259749, "score": 5, "text": "<p>If you just want lots of examples .... note that, if $a$ is rational and not an integer, then\n$$\n\\sum_{n=1}^\\infty\\left(\\frac{1}{n} - \\frac{1}{n+a-1}\\right) = \\gamma+\\psi(a)\n$$\nis pseudo-rational. ($\\psi$ is the digamma)\nI plugged in $k/12$ for $1 \\le k \\le 11$ to get these:\n$$\n-2\\,\\ln \\left( 2 \\right) ,\\\\-1/6\\,\\pi \\,\\sqrt {3}-3/2\\,\\ln \\left( 3\n \\right) ,\\\\-3\\,\\ln \\left( 2 \\right) -\\pi /2,\\\\-3/2\\,\\ln \\left( 3\n \\right) -2\\,\\ln \\left( 2 \\right) -1/2\\,\\pi \\,\\sqrt {3},\\\\-1/2\\,\\sqrt {\n3}\\ln \\left( 2-\\sqrt {3} \\right) -3/2\\,\\ln \\left( 2-\\sqrt {3}\n \\right) -3/4\\,\\sqrt {3}\\ln \\left( 3 \\right) -9/4\\,\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) -1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \n\\pi /12 \\right) -1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \\pi /12 \\right) \n\\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) +3/2\\,\\ln \\left( -\n3+2\\,\\sqrt {3} \\right) \\sqrt {3},\\\\1/6\\,\\pi \\,\\sqrt {3}-3/2\\,\\ln \n \\left( 3 \\right) ,\\\\-3\\,\\ln \\left( 2 \\right) +\\pi /2,\\\\-3/2\\,\\ln \n \\left( 3 \\right) -2\\,\\ln \\left( 2 \\right) +1/2\\,\\pi \\,\\sqrt {3},\\\\-3/2\n\\,\\ln \\left( 2-\\sqrt {3} \\right) +1/2\\,\\sqrt {3}\\ln \\left( 2-\\sqrt {\n3} \\right) -9/4\\,\\ln \\left( 3 \\right) +3/4\\,\\sqrt {3}\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) -1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( {\n\\frac {5\\,\\pi }{12}} \\right) \\sqrt {3}+1/2\\,\\sqrt {2}\\pi \\,\\cos\n \\left( {\\frac {5\\,\\pi }{12}} \\right) -3/2\\,\\ln \\left( -3+2\\,\\sqrt {3\n} \\right) \\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) ,\\\\-3/2\\,\n\\ln \\left( 2-\\sqrt {3} \\right) +1/2\\,\\sqrt {3}\\ln \\left( 2-\\sqrt {3}\n \\right) -9/4\\,\\ln \\left( 3 \\right) +3/4\\,\\sqrt {3}\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) +1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( {\n\\frac {5\\,\\pi }{12}} \\right) \\sqrt {3}-1/2\\,\\sqrt {2}\\pi \\,\\cos\n \\left( {\\frac {5\\,\\pi }{12}} \\right) -3/2\\,\\ln \\left( -3+2\\,\\sqrt {3\n} \\right) \\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) ,\\\\-1/2\\,\n\\sqrt {3}\\ln \\left( 2-\\sqrt {3} \\right) -3/2\\,\\ln \\left( 2-\\sqrt {3}\n \\right) -3/4\\,\\sqrt {3}\\ln \\left( 3 \\right) -9/4\\,\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) +1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \n\\pi /12 \\right) +1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \\pi /12 \\right) \n\\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) +3/2\\,\\ln \\left( -\n3+2\\,\\sqrt {3} \\right) \\sqrt {3}\n$$</p>\n" }, { "answer_id": 259708, "score": 11, "text": "<p><em>One partial answer:</em></p>\n\n<p>$\\pi$ and $\\ln(2)$ are both pseudo-rational:\n\\begin{align}\n\\ln(2) &amp;= \\sum\\frac{1}{2n\\,(2n-1)} \\\\\n\\pi &amp;= \\sum\\frac{3}{n\\,(2n-1)\\,(4n-3)} \\\\\n\\end{align}</p>\n\n<p>These follow from statements in <a href=\"https://en.wikipedia.org/wiki/Digamma_function\">Wikipedia</a>, including Gauss's digamma theorem, and are also asserted by Mathematica. By similar manipulations, $\\pi\\sqrt{3}$ and $\\ln(3)$ are pseudo-rational also.</p>\n\n<p>$\\ $</p>\n\n<p><em>Some questions collected from the comments:</em></p>\n\n<ul>\n<li>Are the pseudo-rationals closed under multiplication?</li>\n<li>Are all pseudo-rationals <a href=\"https://en.wikipedia.org/wiki/Ring_of_periods\">periods</a>?</li>\n<li>Is there a procedure to decide if the pseudo-rational from $P,Q$ is positive? </li>\n</ul>\n\n<p>The last one is difficult because there are non-trivial zeroes like $\\sum\\left(\\dfrac{6}{n^2}-\\dfrac{8}{(2n-1)^2}\\right)$.</p>\n" }, { "answer_id": 259780, "score": 7, "text": "<p>Also possibly of interest:</p>\n\n<p>$$ \\sum_{n=1}^\\infty \\left(-\\frac{2}{b(n-1/b)(n+1/b)} + \\frac{b}{n(n+1)}\\right) = \\pi \\cot(\\pi/b)$$</p>\n\n<p>$$ \\sum_{n=1}^\\infty {\\frac {t \\left( {t}^{2}{n}^{2}+2\\,{n}^{2}+2\\,n+1 \\right) }{ \\left( n+\n1 \\right) n \\left( {t}^{2}{n}^{2}+1 \\right) }}\n = \\pi \\coth(\\pi/t) $$</p>\n\n<p>EDIT: \nAnd, if I'm not mistaken,</p>\n\n<p>$$ \\sum_{n=1}^\\infty \\left(\\frac{1-m}{mn} + \\sum_{k=1}^{m-1} \\frac{1}{mn-k}\\right) = \\ln(m) $$\nfor positive integers $m$.</p>\n" }, { "answer_id": 259749, "score": 5, "text": "<p>If you just want lots of examples .... note that, if $a$ is rational and not an integer, then\n$$\n\\sum_{n=1}^\\infty\\left(\\frac{1}{n} - \\frac{1}{n+a-1}\\right) = \\gamma+\\psi(a)\n$$\nis pseudo-rational. ($\\psi$ is the digamma)\nI plugged in $k/12$ for $1 \\le k \\le 11$ to get these:\n$$\n-2\\,\\ln \\left( 2 \\right) ,\\\\-1/6\\,\\pi \\,\\sqrt {3}-3/2\\,\\ln \\left( 3\n \\right) ,\\\\-3\\,\\ln \\left( 2 \\right) -\\pi /2,\\\\-3/2\\,\\ln \\left( 3\n \\right) -2\\,\\ln \\left( 2 \\right) -1/2\\,\\pi \\,\\sqrt {3},\\\\-1/2\\,\\sqrt {\n3}\\ln \\left( 2-\\sqrt {3} \\right) -3/2\\,\\ln \\left( 2-\\sqrt {3}\n \\right) -3/4\\,\\sqrt {3}\\ln \\left( 3 \\right) -9/4\\,\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) -1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \n\\pi /12 \\right) -1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \\pi /12 \\right) \n\\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) +3/2\\,\\ln \\left( -\n3+2\\,\\sqrt {3} \\right) \\sqrt {3},\\\\1/6\\,\\pi \\,\\sqrt {3}-3/2\\,\\ln \n \\left( 3 \\right) ,\\\\-3\\,\\ln \\left( 2 \\right) +\\pi /2,\\\\-3/2\\,\\ln \n \\left( 3 \\right) -2\\,\\ln \\left( 2 \\right) +1/2\\,\\pi \\,\\sqrt {3},\\\\-3/2\n\\,\\ln \\left( 2-\\sqrt {3} \\right) +1/2\\,\\sqrt {3}\\ln \\left( 2-\\sqrt {\n3} \\right) -9/4\\,\\ln \\left( 3 \\right) +3/4\\,\\sqrt {3}\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) -1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( {\n\\frac {5\\,\\pi }{12}} \\right) \\sqrt {3}+1/2\\,\\sqrt {2}\\pi \\,\\cos\n \\left( {\\frac {5\\,\\pi }{12}} \\right) -3/2\\,\\ln \\left( -3+2\\,\\sqrt {3\n} \\right) \\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) ,\\\\-3/2\\,\n\\ln \\left( 2-\\sqrt {3} \\right) +1/2\\,\\sqrt {3}\\ln \\left( 2-\\sqrt {3}\n \\right) -9/4\\,\\ln \\left( 3 \\right) +3/4\\,\\sqrt {3}\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) +1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( {\n\\frac {5\\,\\pi }{12}} \\right) \\sqrt {3}-1/2\\,\\sqrt {2}\\pi \\,\\cos\n \\left( {\\frac {5\\,\\pi }{12}} \\right) -3/2\\,\\ln \\left( -3+2\\,\\sqrt {3\n} \\right) \\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) ,\\\\-1/2\\,\n\\sqrt {3}\\ln \\left( 2-\\sqrt {3} \\right) -3/2\\,\\ln \\left( 2-\\sqrt {3}\n \\right) -3/4\\,\\sqrt {3}\\ln \\left( 3 \\right) -9/4\\,\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) +1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \n\\pi /12 \\right) +1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \\pi /12 \\right) \n\\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) +3/2\\,\\ln \\left( -\n3+2\\,\\sqrt {3} \\right) \\sqrt {3}\n$$</p>\n" }, { "answer_id": 259708, "score": 11, "text": "<p><em>One partial answer:</em></p>\n\n<p>$\\pi$ and $\\ln(2)$ are both pseudo-rational:\n\\begin{align}\n\\ln(2) &amp;= \\sum\\frac{1}{2n\\,(2n-1)} \\\\\n\\pi &amp;= \\sum\\frac{3}{n\\,(2n-1)\\,(4n-3)} \\\\\n\\end{align}</p>\n\n<p>These follow from statements in <a href=\"https://en.wikipedia.org/wiki/Digamma_function\">Wikipedia</a>, including Gauss's digamma theorem, and are also asserted by Mathematica. By similar manipulations, $\\pi\\sqrt{3}$ and $\\ln(3)$ are pseudo-rational also.</p>\n\n<p>$\\ $</p>\n\n<p><em>Some questions collected from the comments:</em></p>\n\n<ul>\n<li>Are the pseudo-rationals closed under multiplication?</li>\n<li>Are all pseudo-rationals <a href=\"https://en.wikipedia.org/wiki/Ring_of_periods\">periods</a>?</li>\n<li>Is there a procedure to decide if the pseudo-rational from $P,Q$ is positive? </li>\n</ul>\n\n<p>The last one is difficult because there are non-trivial zeroes like $\\sum\\left(\\dfrac{6}{n^2}-\\dfrac{8}{(2n-1)^2}\\right)$.</p>\n" }, { "answer_id": 259780, "score": 7, "text": "<p>Also possibly of interest:</p>\n\n<p>$$ \\sum_{n=1}^\\infty \\left(-\\frac{2}{b(n-1/b)(n+1/b)} + \\frac{b}{n(n+1)}\\right) = \\pi \\cot(\\pi/b)$$</p>\n\n<p>$$ \\sum_{n=1}^\\infty {\\frac {t \\left( {t}^{2}{n}^{2}+2\\,{n}^{2}+2\\,n+1 \\right) }{ \\left( n+\n1 \\right) n \\left( {t}^{2}{n}^{2}+1 \\right) }}\n = \\pi \\coth(\\pi/t) $$</p>\n\n<p>EDIT: \nAnd, if I'm not mistaken,</p>\n\n<p>$$ \\sum_{n=1}^\\infty \\left(\\frac{1-m}{mn} + \\sum_{k=1}^{m-1} \\frac{1}{mn-k}\\right) = \\ln(m) $$\nfor positive integers $m$.</p>\n" }, { "answer_id": 259749, "score": 5, "text": "<p>If you just want lots of examples .... note that, if $a$ is rational and not an integer, then\n$$\n\\sum_{n=1}^\\infty\\left(\\frac{1}{n} - \\frac{1}{n+a-1}\\right) = \\gamma+\\psi(a)\n$$\nis pseudo-rational. ($\\psi$ is the digamma)\nI plugged in $k/12$ for $1 \\le k \\le 11$ to get these:\n$$\n-2\\,\\ln \\left( 2 \\right) ,\\\\-1/6\\,\\pi \\,\\sqrt {3}-3/2\\,\\ln \\left( 3\n \\right) ,\\\\-3\\,\\ln \\left( 2 \\right) -\\pi /2,\\\\-3/2\\,\\ln \\left( 3\n \\right) -2\\,\\ln \\left( 2 \\right) -1/2\\,\\pi \\,\\sqrt {3},\\\\-1/2\\,\\sqrt {\n3}\\ln \\left( 2-\\sqrt {3} \\right) -3/2\\,\\ln \\left( 2-\\sqrt {3}\n \\right) -3/4\\,\\sqrt {3}\\ln \\left( 3 \\right) -9/4\\,\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) -1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \n\\pi /12 \\right) -1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \\pi /12 \\right) \n\\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) +3/2\\,\\ln \\left( -\n3+2\\,\\sqrt {3} \\right) \\sqrt {3},\\\\1/6\\,\\pi \\,\\sqrt {3}-3/2\\,\\ln \n \\left( 3 \\right) ,\\\\-3\\,\\ln \\left( 2 \\right) +\\pi /2,\\\\-3/2\\,\\ln \n \\left( 3 \\right) -2\\,\\ln \\left( 2 \\right) +1/2\\,\\pi \\,\\sqrt {3},\\\\-3/2\n\\,\\ln \\left( 2-\\sqrt {3} \\right) +1/2\\,\\sqrt {3}\\ln \\left( 2-\\sqrt {\n3} \\right) -9/4\\,\\ln \\left( 3 \\right) +3/4\\,\\sqrt {3}\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) -1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( {\n\\frac {5\\,\\pi }{12}} \\right) \\sqrt {3}+1/2\\,\\sqrt {2}\\pi \\,\\cos\n \\left( {\\frac {5\\,\\pi }{12}} \\right) -3/2\\,\\ln \\left( -3+2\\,\\sqrt {3\n} \\right) \\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) ,\\\\-3/2\\,\n\\ln \\left( 2-\\sqrt {3} \\right) +1/2\\,\\sqrt {3}\\ln \\left( 2-\\sqrt {3}\n \\right) -9/4\\,\\ln \\left( 3 \\right) +3/4\\,\\sqrt {3}\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) +1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( {\n\\frac {5\\,\\pi }{12}} \\right) \\sqrt {3}-1/2\\,\\sqrt {2}\\pi \\,\\cos\n \\left( {\\frac {5\\,\\pi }{12}} \\right) -3/2\\,\\ln \\left( -3+2\\,\\sqrt {3\n} \\right) \\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) ,\\\\-1/2\\,\n\\sqrt {3}\\ln \\left( 2-\\sqrt {3} \\right) -3/2\\,\\ln \\left( 2-\\sqrt {3}\n \\right) -3/4\\,\\sqrt {3}\\ln \\left( 3 \\right) -9/4\\,\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) +1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \n\\pi /12 \\right) +1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \\pi /12 \\right) \n\\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) +3/2\\,\\ln \\left( -\n3+2\\,\\sqrt {3} \\right) \\sqrt {3}\n$$</p>\n" }, { "answer_id": 259708, "score": 11, "text": "<p><em>One partial answer:</em></p>\n\n<p>$\\pi$ and $\\ln(2)$ are both pseudo-rational:\n\\begin{align}\n\\ln(2) &amp;= \\sum\\frac{1}{2n\\,(2n-1)} \\\\\n\\pi &amp;= \\sum\\frac{3}{n\\,(2n-1)\\,(4n-3)} \\\\\n\\end{align}</p>\n\n<p>These follow from statements in <a href=\"https://en.wikipedia.org/wiki/Digamma_function\">Wikipedia</a>, including Gauss's digamma theorem, and are also asserted by Mathematica. By similar manipulations, $\\pi\\sqrt{3}$ and $\\ln(3)$ are pseudo-rational also.</p>\n\n<p>$\\ $</p>\n\n<p><em>Some questions collected from the comments:</em></p>\n\n<ul>\n<li>Are the pseudo-rationals closed under multiplication?</li>\n<li>Are all pseudo-rationals <a href=\"https://en.wikipedia.org/wiki/Ring_of_periods\">periods</a>?</li>\n<li>Is there a procedure to decide if the pseudo-rational from $P,Q$ is positive? </li>\n</ul>\n\n<p>The last one is difficult because there are non-trivial zeroes like $\\sum\\left(\\dfrac{6}{n^2}-\\dfrac{8}{(2n-1)^2}\\right)$.</p>\n" }, { "answer_id": 259780, "score": 7, "text": "<p>Also possibly of interest:</p>\n\n<p>$$ \\sum_{n=1}^\\infty \\left(-\\frac{2}{b(n-1/b)(n+1/b)} + \\frac{b}{n(n+1)}\\right) = \\pi \\cot(\\pi/b)$$</p>\n\n<p>$$ \\sum_{n=1}^\\infty {\\frac {t \\left( {t}^{2}{n}^{2}+2\\,{n}^{2}+2\\,n+1 \\right) }{ \\left( n+\n1 \\right) n \\left( {t}^{2}{n}^{2}+1 \\right) }}\n = \\pi \\coth(\\pi/t) $$</p>\n\n<p>EDIT: \nAnd, if I'm not mistaken,</p>\n\n<p>$$ \\sum_{n=1}^\\infty \\left(\\frac{1-m}{mn} + \\sum_{k=1}^{m-1} \\frac{1}{mn-k}\\right) = \\ln(m) $$\nfor positive integers $m$.</p>\n" }, { "answer_id": 259749, "score": 5, "text": "<p>If you just want lots of examples .... note that, if $a$ is rational and not an integer, then\n$$\n\\sum_{n=1}^\\infty\\left(\\frac{1}{n} - \\frac{1}{n+a-1}\\right) = \\gamma+\\psi(a)\n$$\nis pseudo-rational. ($\\psi$ is the digamma)\nI plugged in $k/12$ for $1 \\le k \\le 11$ to get these:\n$$\n-2\\,\\ln \\left( 2 \\right) ,\\\\-1/6\\,\\pi \\,\\sqrt {3}-3/2\\,\\ln \\left( 3\n \\right) ,\\\\-3\\,\\ln \\left( 2 \\right) -\\pi /2,\\\\-3/2\\,\\ln \\left( 3\n \\right) -2\\,\\ln \\left( 2 \\right) -1/2\\,\\pi \\,\\sqrt {3},\\\\-1/2\\,\\sqrt {\n3}\\ln \\left( 2-\\sqrt {3} \\right) -3/2\\,\\ln \\left( 2-\\sqrt {3}\n \\right) -3/4\\,\\sqrt {3}\\ln \\left( 3 \\right) -9/4\\,\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) -1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \n\\pi /12 \\right) -1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \\pi /12 \\right) \n\\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) +3/2\\,\\ln \\left( -\n3+2\\,\\sqrt {3} \\right) \\sqrt {3},\\\\1/6\\,\\pi \\,\\sqrt {3}-3/2\\,\\ln \n \\left( 3 \\right) ,\\\\-3\\,\\ln \\left( 2 \\right) +\\pi /2,\\\\-3/2\\,\\ln \n \\left( 3 \\right) -2\\,\\ln \\left( 2 \\right) +1/2\\,\\pi \\,\\sqrt {3},\\\\-3/2\n\\,\\ln \\left( 2-\\sqrt {3} \\right) +1/2\\,\\sqrt {3}\\ln \\left( 2-\\sqrt {\n3} \\right) -9/4\\,\\ln \\left( 3 \\right) +3/4\\,\\sqrt {3}\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) -1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( {\n\\frac {5\\,\\pi }{12}} \\right) \\sqrt {3}+1/2\\,\\sqrt {2}\\pi \\,\\cos\n \\left( {\\frac {5\\,\\pi }{12}} \\right) -3/2\\,\\ln \\left( -3+2\\,\\sqrt {3\n} \\right) \\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) ,\\\\-3/2\\,\n\\ln \\left( 2-\\sqrt {3} \\right) +1/2\\,\\sqrt {3}\\ln \\left( 2-\\sqrt {3}\n \\right) -9/4\\,\\ln \\left( 3 \\right) +3/4\\,\\sqrt {3}\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) +1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( {\n\\frac {5\\,\\pi }{12}} \\right) \\sqrt {3}-1/2\\,\\sqrt {2}\\pi \\,\\cos\n \\left( {\\frac {5\\,\\pi }{12}} \\right) -3/2\\,\\ln \\left( -3+2\\,\\sqrt {3\n} \\right) \\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) ,\\\\-1/2\\,\n\\sqrt {3}\\ln \\left( 2-\\sqrt {3} \\right) -3/2\\,\\ln \\left( 2-\\sqrt {3}\n \\right) -3/4\\,\\sqrt {3}\\ln \\left( 3 \\right) -9/4\\,\\ln \\left( 3\n \\right) -3\\,\\ln \\left( 2 \\right) +1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \n\\pi /12 \\right) +1/2\\,\\sqrt {2}\\pi \\,\\cos \\left( \\pi /12 \\right) \n\\sqrt {3}+3/2\\,\\ln \\left( -3+2\\,\\sqrt {3} \\right) +3/2\\,\\ln \\left( -\n3+2\\,\\sqrt {3} \\right) \\sqrt {3}\n$$</p>\n" } ]
Question: What is the set of all &quot;pseudo-rational&quot; numbers (see details)? Answer 1: <p><em>One partial answer:</em></p> <p>$\pi$ and $\ln(2)$ are both pseudo-rational: \begin{align} \ln(2) &amp;= \sum\frac{1}{2n\,(2n-1)} \\ \pi &amp;= \sum\frac{3}{n\,(2n-1)\,(4n-3)} \\ \end{align}</p> <p>These follow from statements in <a href="https://en.wikipedia.org/wiki/Digamma_function">Wikipedia</a>, including Gauss's digamma theorem, and are also asserted by Mathematica. By similar manipulations, $\pi\sqrt{3}$ and $\ln(3)$ are pseudo-rational also.</p> <p>$\ $</p> <p><em>Some questions collected from the comments:</em></p> <ul> <li>Are the pseudo-rationals closed under multiplication?</li> <li>Are all pseudo-rationals <a href="https://en.wikipedia.org/wiki/Ring_of_periods">periods</a>?</li> <li>Is there a procedure to decide if the pseudo-rational from $P,Q$ is positive? </li> </ul> <p>The last one is difficult because there are non-trivial zeroes like $\sum\left(\dfrac{6}{n^2}-\dfrac{8}{(2n-1)^2}\right)$.</p> Answer 2: <p>Also possibly of interest:</p> <p>$$ \sum_{n=1}^\infty \left(-\frac{2}{b(n-1/b)(n+1/b)} + \frac{b}{n(n+1)}\right) = \pi \cot(\pi/b)$$</p> <p>$$ \sum_{n=1}^\infty {\frac {t \left( {t}^{2}{n}^{2}+2\,{n}^{2}+2\,n+1 \right) }{ \left( n+ 1 \right) n \left( {t}^{2}{n}^{2}+1 \right) }} = \pi \coth(\pi/t) $$</p> <p>EDIT: And, if I'm not mistaken,</p> <p>$$ \sum_{n=1}^\infty \left(\frac{1-m}{mn} + \sum_{k=1}^{m-1} \frac{1}{mn-k}\right) = \ln(m) $$ for positive integers $m$.</p> Answer 3: <p>If you just want lots of examples .... note that, if $a$ is rational and not an integer, then $$ \sum_{n=1}^\infty\left(\frac{1}{n} - \frac{1}{n+a-1}\right) = \gamma+\psi(a) $$ is pseudo-rational. ($\psi$ is the digamma) I plugged in $k/12$ for $1 \le k \le 11$ to get these: $$ -2\,\ln \left( 2 \right) ,\\-1/6\,\pi \,\sqrt {3}-3/2\,\ln \left( 3 \right) ,\\-3\,\ln \left( 2 \right) -\pi /2,\\-3/2\,\ln \left( 3 \right) -2\,\ln \left( 2 \right) -1/2\,\pi \,\sqrt {3},\\-1/2\,\sqrt { 3}\ln \left( 2-\sqrt {3} \right) -3/2\,\ln \left( 2-\sqrt {3} \right) -3/4\,\sqrt {3}\ln \left( 3 \right) -9/4\,\ln \left( 3 \right) -3\,\ln \left( 2 \right) -1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) -1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) +3/2\,\ln \left( - 3+2\,\sqrt {3} \right) \sqrt {3},\\1/6\,\pi \,\sqrt {3}-3/2\,\ln \left( 3 \right) ,\\-3\,\ln \left( 2 \right) +\pi /2,\\-3/2\,\ln \left( 3 \right) -2\,\ln \left( 2 \right) +1/2\,\pi \,\sqrt {3},\\-3/2 \,\ln \left( 2-\sqrt {3} \right) +1/2\,\sqrt {3}\ln \left( 2-\sqrt { 3} \right) -9/4\,\ln \left( 3 \right) +3/4\,\sqrt {3}\ln \left( 3 \right) -3\,\ln \left( 2 \right) -1/2\,\sqrt {2}\pi \,\cos \left( { \frac {5\,\pi }{12}} \right) \sqrt {3}+1/2\,\sqrt {2}\pi \,\cos \left( {\frac {5\,\pi }{12}} \right) -3/2\,\ln \left( -3+2\,\sqrt {3 } \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) ,\\-3/2\, \ln \left( 2-\sqrt {3} \right) +1/2\,\sqrt {3}\ln \left( 2-\sqrt {3} \right) -9/4\,\ln \left( 3 \right) +3/4\,\sqrt {3}\ln \left( 3 \right) -3\,\ln \left( 2 \right) +1/2\,\sqrt {2}\pi \,\cos \left( { \frac {5\,\pi }{12}} \right) \sqrt {3}-1/2\,\sqrt {2}\pi \,\cos \left( {\frac {5\,\pi }{12}} \right) -3/2\,\ln \left( -3+2\,\sqrt {3 } \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) ,\\-1/2\, \sqrt {3}\ln \left( 2-\sqrt {3} \right) -3/2\,\ln \left( 2-\sqrt {3} \right) -3/4\,\sqrt {3}\ln \left( 3 \right) -9/4\,\ln \left( 3 \right) -3\,\ln \left( 2 \right) +1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) +1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) +3/2\,\ln \left( - 3+2\,\sqrt {3} \right) \sqrt {3} $$</p> Answer 4: <p><em>One partial answer:</em></p> <p>$\pi$ and $\ln(2)$ are both pseudo-rational: \begin{align} \ln(2) &amp;= \sum\frac{1}{2n\,(2n-1)} \\ \pi &amp;= \sum\frac{3}{n\,(2n-1)\,(4n-3)} \\ \end{align}</p> <p>These follow from statements in <a href="https://en.wikipedia.org/wiki/Digamma_function">Wikipedia</a>, including Gauss's digamma theorem, and are also asserted by Mathematica. By similar manipulations, $\pi\sqrt{3}$ and $\ln(3)$ are pseudo-rational also.</p> <p>$\ $</p> <p><em>Some questions collected from the comments:</em></p> <ul> <li>Are the pseudo-rationals closed under multiplication?</li> <li>Are all pseudo-rationals <a href="https://en.wikipedia.org/wiki/Ring_of_periods">periods</a>?</li> <li>Is there a procedure to decide if the pseudo-rational from $P,Q$ is positive? </li> </ul> <p>The last one is difficult because there are non-trivial zeroes like $\sum\left(\dfrac{6}{n^2}-\dfrac{8}{(2n-1)^2}\right)$.</p> Answer 5: <p>Also possibly of interest:</p> <p>$$ \sum_{n=1}^\infty \left(-\frac{2}{b(n-1/b)(n+1/b)} + \frac{b}{n(n+1)}\right) = \pi \cot(\pi/b)$$</p> <p>$$ \sum_{n=1}^\infty {\frac {t \left( {t}^{2}{n}^{2}+2\,{n}^{2}+2\,n+1 \right) }{ \left( n+ 1 \right) n \left( {t}^{2}{n}^{2}+1 \right) }} = \pi \coth(\pi/t) $$</p> <p>EDIT: And, if I'm not mistaken,</p> <p>$$ \sum_{n=1}^\infty \left(\frac{1-m}{mn} + \sum_{k=1}^{m-1} \frac{1}{mn-k}\right) = \ln(m) $$ for positive integers $m$.</p> Answer 6: <p>If you just want lots of examples .... note that, if $a$ is rational and not an integer, then $$ \sum_{n=1}^\infty\left(\frac{1}{n} - \frac{1}{n+a-1}\right) = \gamma+\psi(a) $$ is pseudo-rational. ($\psi$ is the digamma) I plugged in $k/12$ for $1 \le k \le 11$ to get these: $$ -2\,\ln \left( 2 \right) ,\\-1/6\,\pi \,\sqrt {3}-3/2\,\ln \left( 3 \right) ,\\-3\,\ln \left( 2 \right) -\pi /2,\\-3/2\,\ln \left( 3 \right) -2\,\ln \left( 2 \right) -1/2\,\pi \,\sqrt {3},\\-1/2\,\sqrt { 3}\ln \left( 2-\sqrt {3} \right) -3/2\,\ln \left( 2-\sqrt {3} \right) -3/4\,\sqrt {3}\ln \left( 3 \right) -9/4\,\ln \left( 3 \right) -3\,\ln \left( 2 \right) -1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) -1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) +3/2\,\ln \left( - 3+2\,\sqrt {3} \right) \sqrt {3},\\1/6\,\pi \,\sqrt {3}-3/2\,\ln \left( 3 \right) ,\\-3\,\ln \left( 2 \right) +\pi /2,\\-3/2\,\ln \left( 3 \right) -2\,\ln \left( 2 \right) +1/2\,\pi \,\sqrt {3},\\-3/2 \,\ln \left( 2-\sqrt {3} \right) +1/2\,\sqrt {3}\ln \left( 2-\sqrt { 3} \right) -9/4\,\ln \left( 3 \right) +3/4\,\sqrt {3}\ln \left( 3 \right) -3\,\ln \left( 2 \right) -1/2\,\sqrt {2}\pi \,\cos \left( { \frac {5\,\pi }{12}} \right) \sqrt {3}+1/2\,\sqrt {2}\pi \,\cos \left( {\frac {5\,\pi }{12}} \right) -3/2\,\ln \left( -3+2\,\sqrt {3 } \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) ,\\-3/2\, \ln \left( 2-\sqrt {3} \right) +1/2\,\sqrt {3}\ln \left( 2-\sqrt {3} \right) -9/4\,\ln \left( 3 \right) +3/4\,\sqrt {3}\ln \left( 3 \right) -3\,\ln \left( 2 \right) +1/2\,\sqrt {2}\pi \,\cos \left( { \frac {5\,\pi }{12}} \right) \sqrt {3}-1/2\,\sqrt {2}\pi \,\cos \left( {\frac {5\,\pi }{12}} \right) -3/2\,\ln \left( -3+2\,\sqrt {3 } \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) ,\\-1/2\, \sqrt {3}\ln \left( 2-\sqrt {3} \right) -3/2\,\ln \left( 2-\sqrt {3} \right) -3/4\,\sqrt {3}\ln \left( 3 \right) -9/4\,\ln \left( 3 \right) -3\,\ln \left( 2 \right) +1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) +1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) +3/2\,\ln \left( - 3+2\,\sqrt {3} \right) \sqrt {3} $$</p> Answer 7: <p><em>One partial answer:</em></p> <p>$\pi$ and $\ln(2)$ are both pseudo-rational: \begin{align} \ln(2) &amp;= \sum\frac{1}{2n\,(2n-1)} \\ \pi &amp;= \sum\frac{3}{n\,(2n-1)\,(4n-3)} \\ \end{align}</p> <p>These follow from statements in <a href="https://en.wikipedia.org/wiki/Digamma_function">Wikipedia</a>, including Gauss's digamma theorem, and are also asserted by Mathematica. By similar manipulations, $\pi\sqrt{3}$ and $\ln(3)$ are pseudo-rational also.</p> <p>$\ $</p> <p><em>Some questions collected from the comments:</em></p> <ul> <li>Are the pseudo-rationals closed under multiplication?</li> <li>Are all pseudo-rationals <a href="https://en.wikipedia.org/wiki/Ring_of_periods">periods</a>?</li> <li>Is there a procedure to decide if the pseudo-rational from $P,Q$ is positive? </li> </ul> <p>The last one is difficult because there are non-trivial zeroes like $\sum\left(\dfrac{6}{n^2}-\dfrac{8}{(2n-1)^2}\right)$.</p> Answer 8: <p>Also possibly of interest:</p> <p>$$ \sum_{n=1}^\infty \left(-\frac{2}{b(n-1/b)(n+1/b)} + \frac{b}{n(n+1)}\right) = \pi \cot(\pi/b)$$</p> <p>$$ \sum_{n=1}^\infty {\frac {t \left( {t}^{2}{n}^{2}+2\,{n}^{2}+2\,n+1 \right) }{ \left( n+ 1 \right) n \left( {t}^{2}{n}^{2}+1 \right) }} = \pi \coth(\pi/t) $$</p> <p>EDIT: And, if I'm not mistaken,</p> <p>$$ \sum_{n=1}^\infty \left(\frac{1-m}{mn} + \sum_{k=1}^{m-1} \frac{1}{mn-k}\right) = \ln(m) $$ for positive integers $m$.</p> Answer 9: <p>If you just want lots of examples .... note that, if $a$ is rational and not an integer, then $$ \sum_{n=1}^\infty\left(\frac{1}{n} - \frac{1}{n+a-1}\right) = \gamma+\psi(a) $$ is pseudo-rational. ($\psi$ is the digamma) I plugged in $k/12$ for $1 \le k \le 11$ to get these: $$ -2\,\ln \left( 2 \right) ,\\-1/6\,\pi \,\sqrt {3}-3/2\,\ln \left( 3 \right) ,\\-3\,\ln \left( 2 \right) -\pi /2,\\-3/2\,\ln \left( 3 \right) -2\,\ln \left( 2 \right) -1/2\,\pi \,\sqrt {3},\\-1/2\,\sqrt { 3}\ln \left( 2-\sqrt {3} \right) -3/2\,\ln \left( 2-\sqrt {3} \right) -3/4\,\sqrt {3}\ln \left( 3 \right) -9/4\,\ln \left( 3 \right) -3\,\ln \left( 2 \right) -1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) -1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) +3/2\,\ln \left( - 3+2\,\sqrt {3} \right) \sqrt {3},\\1/6\,\pi \,\sqrt {3}-3/2\,\ln \left( 3 \right) ,\\-3\,\ln \left( 2 \right) +\pi /2,\\-3/2\,\ln \left( 3 \right) -2\,\ln \left( 2 \right) +1/2\,\pi \,\sqrt {3},\\-3/2 \,\ln \left( 2-\sqrt {3} \right) +1/2\,\sqrt {3}\ln \left( 2-\sqrt { 3} \right) -9/4\,\ln \left( 3 \right) +3/4\,\sqrt {3}\ln \left( 3 \right) -3\,\ln \left( 2 \right) -1/2\,\sqrt {2}\pi \,\cos \left( { \frac {5\,\pi }{12}} \right) \sqrt {3}+1/2\,\sqrt {2}\pi \,\cos \left( {\frac {5\,\pi }{12}} \right) -3/2\,\ln \left( -3+2\,\sqrt {3 } \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) ,\\-3/2\, \ln \left( 2-\sqrt {3} \right) +1/2\,\sqrt {3}\ln \left( 2-\sqrt {3} \right) -9/4\,\ln \left( 3 \right) +3/4\,\sqrt {3}\ln \left( 3 \right) -3\,\ln \left( 2 \right) +1/2\,\sqrt {2}\pi \,\cos \left( { \frac {5\,\pi }{12}} \right) \sqrt {3}-1/2\,\sqrt {2}\pi \,\cos \left( {\frac {5\,\pi }{12}} \right) -3/2\,\ln \left( -3+2\,\sqrt {3 } \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) ,\\-1/2\, \sqrt {3}\ln \left( 2-\sqrt {3} \right) -3/2\,\ln \left( 2-\sqrt {3} \right) -3/4\,\sqrt {3}\ln \left( 3 \right) -9/4\,\ln \left( 3 \right) -3\,\ln \left( 2 \right) +1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) +1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) +3/2\,\ln \left( - 3+2\,\sqrt {3} \right) \sqrt {3} $$</p> Answer 10: <p><em>One partial answer:</em></p> <p>$\pi$ and $\ln(2)$ are both pseudo-rational: \begin{align} \ln(2) &amp;= \sum\frac{1}{2n\,(2n-1)} \\ \pi &amp;= \sum\frac{3}{n\,(2n-1)\,(4n-3)} \\ \end{align}</p> <p>These follow from statements in <a href="https://en.wikipedia.org/wiki/Digamma_function">Wikipedia</a>, including Gauss's digamma theorem, and are also asserted by Mathematica. By similar manipulations, $\pi\sqrt{3}$ and $\ln(3)$ are pseudo-rational also.</p> <p>$\ $</p> <p><em>Some questions collected from the comments:</em></p> <ul> <li>Are the pseudo-rationals closed under multiplication?</li> <li>Are all pseudo-rationals <a href="https://en.wikipedia.org/wiki/Ring_of_periods">periods</a>?</li> <li>Is there a procedure to decide if the pseudo-rational from $P,Q$ is positive? </li> </ul> <p>The last one is difficult because there are non-trivial zeroes like $\sum\left(\dfrac{6}{n^2}-\dfrac{8}{(2n-1)^2}\right)$.</p> Answer 11: <p>Also possibly of interest:</p> <p>$$ \sum_{n=1}^\infty \left(-\frac{2}{b(n-1/b)(n+1/b)} + \frac{b}{n(n+1)}\right) = \pi \cot(\pi/b)$$</p> <p>$$ \sum_{n=1}^\infty {\frac {t \left( {t}^{2}{n}^{2}+2\,{n}^{2}+2\,n+1 \right) }{ \left( n+ 1 \right) n \left( {t}^{2}{n}^{2}+1 \right) }} = \pi \coth(\pi/t) $$</p> <p>EDIT: And, if I'm not mistaken,</p> <p>$$ \sum_{n=1}^\infty \left(\frac{1-m}{mn} + \sum_{k=1}^{m-1} \frac{1}{mn-k}\right) = \ln(m) $$ for positive integers $m$.</p> Answer 12: <p>If you just want lots of examples .... note that, if $a$ is rational and not an integer, then $$ \sum_{n=1}^\infty\left(\frac{1}{n} - \frac{1}{n+a-1}\right) = \gamma+\psi(a) $$ is pseudo-rational. ($\psi$ is the digamma) I plugged in $k/12$ for $1 \le k \le 11$ to get these: $$ -2\,\ln \left( 2 \right) ,\\-1/6\,\pi \,\sqrt {3}-3/2\,\ln \left( 3 \right) ,\\-3\,\ln \left( 2 \right) -\pi /2,\\-3/2\,\ln \left( 3 \right) -2\,\ln \left( 2 \right) -1/2\,\pi \,\sqrt {3},\\-1/2\,\sqrt { 3}\ln \left( 2-\sqrt {3} \right) -3/2\,\ln \left( 2-\sqrt {3} \right) -3/4\,\sqrt {3}\ln \left( 3 \right) -9/4\,\ln \left( 3 \right) -3\,\ln \left( 2 \right) -1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) -1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) +3/2\,\ln \left( - 3+2\,\sqrt {3} \right) \sqrt {3},\\1/6\,\pi \,\sqrt {3}-3/2\,\ln \left( 3 \right) ,\\-3\,\ln \left( 2 \right) +\pi /2,\\-3/2\,\ln \left( 3 \right) -2\,\ln \left( 2 \right) +1/2\,\pi \,\sqrt {3},\\-3/2 \,\ln \left( 2-\sqrt {3} \right) +1/2\,\sqrt {3}\ln \left( 2-\sqrt { 3} \right) -9/4\,\ln \left( 3 \right) +3/4\,\sqrt {3}\ln \left( 3 \right) -3\,\ln \left( 2 \right) -1/2\,\sqrt {2}\pi \,\cos \left( { \frac {5\,\pi }{12}} \right) \sqrt {3}+1/2\,\sqrt {2}\pi \,\cos \left( {\frac {5\,\pi }{12}} \right) -3/2\,\ln \left( -3+2\,\sqrt {3 } \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) ,\\-3/2\, \ln \left( 2-\sqrt {3} \right) +1/2\,\sqrt {3}\ln \left( 2-\sqrt {3} \right) -9/4\,\ln \left( 3 \right) +3/4\,\sqrt {3}\ln \left( 3 \right) -3\,\ln \left( 2 \right) +1/2\,\sqrt {2}\pi \,\cos \left( { \frac {5\,\pi }{12}} \right) \sqrt {3}-1/2\,\sqrt {2}\pi \,\cos \left( {\frac {5\,\pi }{12}} \right) -3/2\,\ln \left( -3+2\,\sqrt {3 } \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) ,\\-1/2\, \sqrt {3}\ln \left( 2-\sqrt {3} \right) -3/2\,\ln \left( 2-\sqrt {3} \right) -3/4\,\sqrt {3}\ln \left( 3 \right) -9/4\,\ln \left( 3 \right) -3\,\ln \left( 2 \right) +1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) +1/2\,\sqrt {2}\pi \,\cos \left( \pi /12 \right) \sqrt {3}+3/2\,\ln \left( -3+2\,\sqrt {3} \right) +3/2\,\ln \left( - 3+2\,\sqrt {3} \right) \sqrt {3} $$</p>
Where does one go to learn about DG-algebras?
260,434
21
https://mathoverflow.net/questions/260434/where-does-one-go-to-learn-about-dg-algebras
<p>The theory of differential graded algebras (in char 0) and their modules has numerous applications in rational homotopy theory as well as algebraic geometry.</p> <p>I'm looking for a reasonably complete reference book/collection of articles that treat the basic algebraic and homotopical aspects of dg-algebras akin to Atiyah-MacDonald's book on commutative algebra. </p> <p>Hopefully containing the following topics</p> <ol> <li>Model structures on dg algebras and simplicial algebras (presenting the "correct" infinity category) and relation between them (concrete description of fibrations and cofibrations).</li> <li>Resolutions and minimal models</li> <li>Localizations and Completions</li> <li>Analogs of the classical types of morphisms (flat, smooth, unramified, etale, open immersion, closed embedding, finite, finite type, etc.).</li> <li>Derived categories of modules (Triangulated / Stable $\infty$) and functors on them (push-pull-shriek functors, perfect complexes vs. bounded vs. unbounded derived categories, dualizing complex, cotangent complex, localization and completion, fourier mukai transforms).</li> <li>Derived descent theorems - descent for dg-modules in the derived categories. (hopefully making it clear what kind of information goes into glueing a dg-module).</li> <li>Hochshild and Cyclic (co-)homology</li> <li>DG lie algebras and Koszul duality</li> </ol> <p>I realize that one can always go to Lurie's books for a comprehensive and much more general treatment but I hope that a more elementary reference exists. I do believe that some aspects of this theory are a lot less technical than they appear.</p>
[ { "answer_id": 261388, "score": 12, "text": "<p>I am not aware of a single book that has all of these topics, but I can list a couple of very good books that can do the job together. First, Benoit Fresse works in this setting quite a bit. His book <a href=\"http://math.univ-lille1.fr/~fresse/OperadModuleFunctors-Updated.pdf\" rel=\"noreferrer\">Modules over Operads and Functors</a> contains material on (1), (2), (5), (7), and (8). The model categorical work is very explicit, including a description of (co)fibrations as you want. I also think I learned the most about Koszul duality from this source. It does not focus on the triangulated structure for (5). A good reference for that is the paper <a href=\"http://homepages.math.uic.edu/~bshipley/classTopFinal.pdf\" rel=\"noreferrer\">Stable Model Categories are Categories of Modules</a>. Fresse also does not focus a ton on (2), but Loday and Vallette do in their book <a href=\"http://www-irma.u-strasbg.fr/~loday/PAPERS/LodayVallette.pdf\" rel=\"noreferrer\">Algebraic Operads</a>. This is also a great source for (1), (7), and (8). The model categorical work goes back to Hinich.</p>\n\n<p>For (3), I recommend <a href=\"https://www.math.uchicago.edu/~may/TEAK/KateBookFinal.pdf\" rel=\"noreferrer\">More Concise Algebraic Topology</a>, by May and Ponto. They also have a discussion of (1).</p>\n\n<p>For (4) and (6), and the more algebraic parts of (5) (e.g. Fourier-Mukai), I recommend <a href=\"https://books.google.com/books?id=vptrCQAAQBAJ&amp;printsec=frontcover\" rel=\"noreferrer\">Axiomatic, Enriched and Motivic Homotopy Theory</a>, edited by John Greenlees, especially Strickland's article about axiomatic homotopy theory. Related is the monograph <a href=\"https://pdfs.semanticscholar.org/55cb/4a34afd7caeb2b4dfb1a08c3c70aee28a9af.pdf\" rel=\"noreferrer\">Axiomatic Stable Homotopy Theory</a>, by Hovey, Palmieri, Strickland. I confess, this is the part of your question I know the least about. Strickland, however, has written in these sources about the connection between the items you list and homotopy theory. Others may have written more on the subject (e.g. <a href=\"http://imperium.lenin.ru/~kaledin/seoul/seoul1-2.pdf\" rel=\"noreferrer\">Kaledin</a>), and I hope someone can come along and add a comment if they know of a source.</p>\n" }, { "answer_id": 316183, "score": 9, "text": "<p>There is a study of DG algebra (DG rings, DG modules, DG categories, DG functors) in the book below. With the related derived categories, etc.</p>\n\n<ul>\n<li><a href=\"https://arxiv.org/abs/1610.09640\" rel=\"noreferrer\">Derived Categories</a></li>\n</ul>\n" }, { "answer_id": 261383, "score": 2, "text": "<p>The <A HREF=\"http://stacks.math.columbia.edu/about\" rel=\"nofollow noreferrer\">Stacks Project</A> at Columbia University hosts a colloborative web-based <A HREF=\"http://stacks.math.columbia.edu/browse\" rel=\"nofollow noreferrer\">survey</A> on algebraic stacks, including a survey of <A HREF=\"http://stacks.math.columbia.edu/chapter/22\" rel=\"nofollow noreferrer\"> Differential Graded Algebras.</A> The whole project is quite an impressive resource.</p>\n\n<p>There is an <A HREF=\"http://math.stackexchange.com/questions/1797368/where-can-i-learn-about-differential-graded-algebras\">MSE posting</A> from last year that lists a variety of sources that target more specific aspects, including:\n<A HREF=\"http://www.maths.ed.ac.uk/~mbooth/hodgeproject.pdf\" rel=\"nofollow noreferrer\">Differential Graded Algebras and Applications</A>, for DG Lie algebras and Koszul duality.</p>\n" }, { "answer_id": 261388, "score": 12, "text": "<p>I am not aware of a single book that has all of these topics, but I can list a couple of very good books that can do the job together. First, Benoit Fresse works in this setting quite a bit. His book <a href=\"http://math.univ-lille1.fr/~fresse/OperadModuleFunctors-Updated.pdf\" rel=\"noreferrer\">Modules over Operads and Functors</a> contains material on (1), (2), (5), (7), and (8). The model categorical work is very explicit, including a description of (co)fibrations as you want. I also think I learned the most about Koszul duality from this source. It does not focus on the triangulated structure for (5). A good reference for that is the paper <a href=\"http://homepages.math.uic.edu/~bshipley/classTopFinal.pdf\" rel=\"noreferrer\">Stable Model Categories are Categories of Modules</a>. Fresse also does not focus a ton on (2), but Loday and Vallette do in their book <a href=\"http://www-irma.u-strasbg.fr/~loday/PAPERS/LodayVallette.pdf\" rel=\"noreferrer\">Algebraic Operads</a>. This is also a great source for (1), (7), and (8). The model categorical work goes back to Hinich.</p>\n\n<p>For (3), I recommend <a href=\"https://www.math.uchicago.edu/~may/TEAK/KateBookFinal.pdf\" rel=\"noreferrer\">More Concise Algebraic Topology</a>, by May and Ponto. They also have a discussion of (1).</p>\n\n<p>For (4) and (6), and the more algebraic parts of (5) (e.g. Fourier-Mukai), I recommend <a href=\"https://books.google.com/books?id=vptrCQAAQBAJ&amp;printsec=frontcover\" rel=\"noreferrer\">Axiomatic, Enriched and Motivic Homotopy Theory</a>, edited by John Greenlees, especially Strickland's article about axiomatic homotopy theory. Related is the monograph <a href=\"https://pdfs.semanticscholar.org/55cb/4a34afd7caeb2b4dfb1a08c3c70aee28a9af.pdf\" rel=\"noreferrer\">Axiomatic Stable Homotopy Theory</a>, by Hovey, Palmieri, Strickland. I confess, this is the part of your question I know the least about. Strickland, however, has written in these sources about the connection between the items you list and homotopy theory. Others may have written more on the subject (e.g. <a href=\"http://imperium.lenin.ru/~kaledin/seoul/seoul1-2.pdf\" rel=\"noreferrer\">Kaledin</a>), and I hope someone can come along and add a comment if they know of a source.</p>\n" }, { "answer_id": 316183, "score": 9, "text": "<p>There is a study of DG algebra (DG rings, DG modules, DG categories, DG functors) in the book below. With the related derived categories, etc.</p>\n\n<ul>\n<li><a href=\"https://arxiv.org/abs/1610.09640\" rel=\"noreferrer\">Derived Categories</a></li>\n</ul>\n" }, { "answer_id": 261383, "score": 2, "text": "<p>The <A HREF=\"http://stacks.math.columbia.edu/about\" rel=\"nofollow noreferrer\">Stacks Project</A> at Columbia University hosts a colloborative web-based <A HREF=\"http://stacks.math.columbia.edu/browse\" rel=\"nofollow noreferrer\">survey</A> on algebraic stacks, including a survey of <A HREF=\"http://stacks.math.columbia.edu/chapter/22\" rel=\"nofollow noreferrer\"> Differential Graded Algebras.</A> The whole project is quite an impressive resource.</p>\n\n<p>There is an <A HREF=\"http://math.stackexchange.com/questions/1797368/where-can-i-learn-about-differential-graded-algebras\">MSE posting</A> from last year that lists a variety of sources that target more specific aspects, including:\n<A HREF=\"http://www.maths.ed.ac.uk/~mbooth/hodgeproject.pdf\" rel=\"nofollow noreferrer\">Differential Graded Algebras and Applications</A>, for DG Lie algebras and Koszul duality.</p>\n" }, { "answer_id": 261388, "score": 12, "text": "<p>I am not aware of a single book that has all of these topics, but I can list a couple of very good books that can do the job together. First, Benoit Fresse works in this setting quite a bit. His book <a href=\"http://math.univ-lille1.fr/~fresse/OperadModuleFunctors-Updated.pdf\" rel=\"noreferrer\">Modules over Operads and Functors</a> contains material on (1), (2), (5), (7), and (8). The model categorical work is very explicit, including a description of (co)fibrations as you want. I also think I learned the most about Koszul duality from this source. It does not focus on the triangulated structure for (5). A good reference for that is the paper <a href=\"http://homepages.math.uic.edu/~bshipley/classTopFinal.pdf\" rel=\"noreferrer\">Stable Model Categories are Categories of Modules</a>. Fresse also does not focus a ton on (2), but Loday and Vallette do in their book <a href=\"http://www-irma.u-strasbg.fr/~loday/PAPERS/LodayVallette.pdf\" rel=\"noreferrer\">Algebraic Operads</a>. This is also a great source for (1), (7), and (8). The model categorical work goes back to Hinich.</p>\n\n<p>For (3), I recommend <a href=\"https://www.math.uchicago.edu/~may/TEAK/KateBookFinal.pdf\" rel=\"noreferrer\">More Concise Algebraic Topology</a>, by May and Ponto. They also have a discussion of (1).</p>\n\n<p>For (4) and (6), and the more algebraic parts of (5) (e.g. Fourier-Mukai), I recommend <a href=\"https://books.google.com/books?id=vptrCQAAQBAJ&amp;printsec=frontcover\" rel=\"noreferrer\">Axiomatic, Enriched and Motivic Homotopy Theory</a>, edited by John Greenlees, especially Strickland's article about axiomatic homotopy theory. Related is the monograph <a href=\"https://pdfs.semanticscholar.org/55cb/4a34afd7caeb2b4dfb1a08c3c70aee28a9af.pdf\" rel=\"noreferrer\">Axiomatic Stable Homotopy Theory</a>, by Hovey, Palmieri, Strickland. I confess, this is the part of your question I know the least about. Strickland, however, has written in these sources about the connection between the items you list and homotopy theory. Others may have written more on the subject (e.g. <a href=\"http://imperium.lenin.ru/~kaledin/seoul/seoul1-2.pdf\" rel=\"noreferrer\">Kaledin</a>), and I hope someone can come along and add a comment if they know of a source.</p>\n" }, { "answer_id": 316183, "score": 9, "text": "<p>There is a study of DG algebra (DG rings, DG modules, DG categories, DG functors) in the book below. With the related derived categories, etc.</p>\n\n<ul>\n<li><a href=\"https://arxiv.org/abs/1610.09640\" rel=\"noreferrer\">Derived Categories</a></li>\n</ul>\n" }, { "answer_id": 261383, "score": 2, "text": "<p>The <A HREF=\"http://stacks.math.columbia.edu/about\" rel=\"nofollow noreferrer\">Stacks Project</A> at Columbia University hosts a colloborative web-based <A HREF=\"http://stacks.math.columbia.edu/browse\" rel=\"nofollow noreferrer\">survey</A> on algebraic stacks, including a survey of <A HREF=\"http://stacks.math.columbia.edu/chapter/22\" rel=\"nofollow noreferrer\"> Differential Graded Algebras.</A> The whole project is quite an impressive resource.</p>\n\n<p>There is an <A HREF=\"http://math.stackexchange.com/questions/1797368/where-can-i-learn-about-differential-graded-algebras\">MSE posting</A> from last year that lists a variety of sources that target more specific aspects, including:\n<A HREF=\"http://www.maths.ed.ac.uk/~mbooth/hodgeproject.pdf\" rel=\"nofollow noreferrer\">Differential Graded Algebras and Applications</A>, for DG Lie algebras and Koszul duality.</p>\n" } ]
Question: Where does one go to learn about DG-algebras? Answer 1: <p>I am not aware of a single book that has all of these topics, but I can list a couple of very good books that can do the job together. First, Benoit Fresse works in this setting quite a bit. His book <a href="http://math.univ-lille1.fr/~fresse/OperadModuleFunctors-Updated.pdf" rel="noreferrer">Modules over Operads and Functors</a> contains material on (1), (2), (5), (7), and (8). The model categorical work is very explicit, including a description of (co)fibrations as you want. I also think I learned the most about Koszul duality from this source. It does not focus on the triangulated structure for (5). A good reference for that is the paper <a href="http://homepages.math.uic.edu/~bshipley/classTopFinal.pdf" rel="noreferrer">Stable Model Categories are Categories of Modules</a>. Fresse also does not focus a ton on (2), but Loday and Vallette do in their book <a href="http://www-irma.u-strasbg.fr/~loday/PAPERS/LodayVallette.pdf" rel="noreferrer">Algebraic Operads</a>. This is also a great source for (1), (7), and (8). The model categorical work goes back to Hinich.</p> <p>For (3), I recommend <a href="https://www.math.uchicago.edu/~may/TEAK/KateBookFinal.pdf" rel="noreferrer">More Concise Algebraic Topology</a>, by May and Ponto. They also have a discussion of (1).</p> <p>For (4) and (6), and the more algebraic parts of (5) (e.g. Fourier-Mukai), I recommend <a href="https://books.google.com/books?id=vptrCQAAQBAJ&amp;printsec=frontcover" rel="noreferrer">Axiomatic, Enriched and Motivic Homotopy Theory</a>, edited by John Greenlees, especially Strickland's article about axiomatic homotopy theory. Related is the monograph <a href="https://pdfs.semanticscholar.org/55cb/4a34afd7caeb2b4dfb1a08c3c70aee28a9af.pdf" rel="noreferrer">Axiomatic Stable Homotopy Theory</a>, by Hovey, Palmieri, Strickland. I confess, this is the part of your question I know the least about. Strickland, however, has written in these sources about the connection between the items you list and homotopy theory. Others may have written more on the subject (e.g. <a href="http://imperium.lenin.ru/~kaledin/seoul/seoul1-2.pdf" rel="noreferrer">Kaledin</a>), and I hope someone can come along and add a comment if they know of a source.</p> Answer 2: <p>There is a study of DG algebra (DG rings, DG modules, DG categories, DG functors) in the book below. With the related derived categories, etc.</p> <ul> <li><a href="https://arxiv.org/abs/1610.09640" rel="noreferrer">Derived Categories</a></li> </ul> Answer 3: <p>The <A HREF="http://stacks.math.columbia.edu/about" rel="nofollow noreferrer">Stacks Project</A> at Columbia University hosts a colloborative web-based <A HREF="http://stacks.math.columbia.edu/browse" rel="nofollow noreferrer">survey</A> on algebraic stacks, including a survey of <A HREF="http://stacks.math.columbia.edu/chapter/22" rel="nofollow noreferrer"> Differential Graded Algebras.</A> The whole project is quite an impressive resource.</p> <p>There is an <A HREF="http://math.stackexchange.com/questions/1797368/where-can-i-learn-about-differential-graded-algebras">MSE posting</A> from last year that lists a variety of sources that target more specific aspects, including: <A HREF="http://www.maths.ed.ac.uk/~mbooth/hodgeproject.pdf" rel="nofollow noreferrer">Differential Graded Algebras and Applications</A>, for DG Lie algebras and Koszul duality.</p> Answer 4: <p>I am not aware of a single book that has all of these topics, but I can list a couple of very good books that can do the job together. First, Benoit Fresse works in this setting quite a bit. His book <a href="http://math.univ-lille1.fr/~fresse/OperadModuleFunctors-Updated.pdf" rel="noreferrer">Modules over Operads and Functors</a> contains material on (1), (2), (5), (7), and (8). The model categorical work is very explicit, including a description of (co)fibrations as you want. I also think I learned the most about Koszul duality from this source. It does not focus on the triangulated structure for (5). A good reference for that is the paper <a href="http://homepages.math.uic.edu/~bshipley/classTopFinal.pdf" rel="noreferrer">Stable Model Categories are Categories of Modules</a>. Fresse also does not focus a ton on (2), but Loday and Vallette do in their book <a href="http://www-irma.u-strasbg.fr/~loday/PAPERS/LodayVallette.pdf" rel="noreferrer">Algebraic Operads</a>. This is also a great source for (1), (7), and (8). The model categorical work goes back to Hinich.</p> <p>For (3), I recommend <a href="https://www.math.uchicago.edu/~may/TEAK/KateBookFinal.pdf" rel="noreferrer">More Concise Algebraic Topology</a>, by May and Ponto. They also have a discussion of (1).</p> <p>For (4) and (6), and the more algebraic parts of (5) (e.g. Fourier-Mukai), I recommend <a href="https://books.google.com/books?id=vptrCQAAQBAJ&amp;printsec=frontcover" rel="noreferrer">Axiomatic, Enriched and Motivic Homotopy Theory</a>, edited by John Greenlees, especially Strickland's article about axiomatic homotopy theory. Related is the monograph <a href="https://pdfs.semanticscholar.org/55cb/4a34afd7caeb2b4dfb1a08c3c70aee28a9af.pdf" rel="noreferrer">Axiomatic Stable Homotopy Theory</a>, by Hovey, Palmieri, Strickland. I confess, this is the part of your question I know the least about. Strickland, however, has written in these sources about the connection between the items you list and homotopy theory. Others may have written more on the subject (e.g. <a href="http://imperium.lenin.ru/~kaledin/seoul/seoul1-2.pdf" rel="noreferrer">Kaledin</a>), and I hope someone can come along and add a comment if they know of a source.</p> Answer 5: <p>There is a study of DG algebra (DG rings, DG modules, DG categories, DG functors) in the book below. With the related derived categories, etc.</p> <ul> <li><a href="https://arxiv.org/abs/1610.09640" rel="noreferrer">Derived Categories</a></li> </ul> Answer 6: <p>The <A HREF="http://stacks.math.columbia.edu/about" rel="nofollow noreferrer">Stacks Project</A> at Columbia University hosts a colloborative web-based <A HREF="http://stacks.math.columbia.edu/browse" rel="nofollow noreferrer">survey</A> on algebraic stacks, including a survey of <A HREF="http://stacks.math.columbia.edu/chapter/22" rel="nofollow noreferrer"> Differential Graded Algebras.</A> The whole project is quite an impressive resource.</p> <p>There is an <A HREF="http://math.stackexchange.com/questions/1797368/where-can-i-learn-about-differential-graded-algebras">MSE posting</A> from last year that lists a variety of sources that target more specific aspects, including: <A HREF="http://www.maths.ed.ac.uk/~mbooth/hodgeproject.pdf" rel="nofollow noreferrer">Differential Graded Algebras and Applications</A>, for DG Lie algebras and Koszul duality.</p> Answer 7: <p>I am not aware of a single book that has all of these topics, but I can list a couple of very good books that can do the job together. First, Benoit Fresse works in this setting quite a bit. His book <a href="http://math.univ-lille1.fr/~fresse/OperadModuleFunctors-Updated.pdf" rel="noreferrer">Modules over Operads and Functors</a> contains material on (1), (2), (5), (7), and (8). The model categorical work is very explicit, including a description of (co)fibrations as you want. I also think I learned the most about Koszul duality from this source. It does not focus on the triangulated structure for (5). A good reference for that is the paper <a href="http://homepages.math.uic.edu/~bshipley/classTopFinal.pdf" rel="noreferrer">Stable Model Categories are Categories of Modules</a>. Fresse also does not focus a ton on (2), but Loday and Vallette do in their book <a href="http://www-irma.u-strasbg.fr/~loday/PAPERS/LodayVallette.pdf" rel="noreferrer">Algebraic Operads</a>. This is also a great source for (1), (7), and (8). The model categorical work goes back to Hinich.</p> <p>For (3), I recommend <a href="https://www.math.uchicago.edu/~may/TEAK/KateBookFinal.pdf" rel="noreferrer">More Concise Algebraic Topology</a>, by May and Ponto. They also have a discussion of (1).</p> <p>For (4) and (6), and the more algebraic parts of (5) (e.g. Fourier-Mukai), I recommend <a href="https://books.google.com/books?id=vptrCQAAQBAJ&amp;printsec=frontcover" rel="noreferrer">Axiomatic, Enriched and Motivic Homotopy Theory</a>, edited by John Greenlees, especially Strickland's article about axiomatic homotopy theory. Related is the monograph <a href="https://pdfs.semanticscholar.org/55cb/4a34afd7caeb2b4dfb1a08c3c70aee28a9af.pdf" rel="noreferrer">Axiomatic Stable Homotopy Theory</a>, by Hovey, Palmieri, Strickland. I confess, this is the part of your question I know the least about. Strickland, however, has written in these sources about the connection between the items you list and homotopy theory. Others may have written more on the subject (e.g. <a href="http://imperium.lenin.ru/~kaledin/seoul/seoul1-2.pdf" rel="noreferrer">Kaledin</a>), and I hope someone can come along and add a comment if they know of a source.</p> Answer 8: <p>There is a study of DG algebra (DG rings, DG modules, DG categories, DG functors) in the book below. With the related derived categories, etc.</p> <ul> <li><a href="https://arxiv.org/abs/1610.09640" rel="noreferrer">Derived Categories</a></li> </ul> Answer 9: <p>The <A HREF="http://stacks.math.columbia.edu/about" rel="nofollow noreferrer">Stacks Project</A> at Columbia University hosts a colloborative web-based <A HREF="http://stacks.math.columbia.edu/browse" rel="nofollow noreferrer">survey</A> on algebraic stacks, including a survey of <A HREF="http://stacks.math.columbia.edu/chapter/22" rel="nofollow noreferrer"> Differential Graded Algebras.</A> The whole project is quite an impressive resource.</p> <p>There is an <A HREF="http://math.stackexchange.com/questions/1797368/where-can-i-learn-about-differential-graded-algebras">MSE posting</A> from last year that lists a variety of sources that target more specific aspects, including: <A HREF="http://www.maths.ed.ac.uk/~mbooth/hodgeproject.pdf" rel="nofollow noreferrer">Differential Graded Algebras and Applications</A>, for DG Lie algebras and Koszul duality.</p>
Theorems that led to very successful research programs in Geometry and Topology
258,818
19
https://mathoverflow.net/questions/258818/theorems-that-led-to-very-successful-research-programs-in-geometry-and-topology
<p>In the recent times I have heard a lot about the following:</p> <ol> <li>The Atiyah-Singer Index theorem</li> <li>H-principle of Gromov ( and others )</li> </ol> <p>It seems to me that these results led to decades of successful research in terms of improving the result, giving a simpler proof or new applications. </p> <p>This motivates me to ask: What are the other such results which led to very successful research programs? I would add Bott Periodicity also.</p> <p>Edit: With my thin background in geometry and topology I cannot make the question more precise or add more examples but I can tell why I would like to know such programs. After I heard these two theorems and their applications I thought every mathematician should know a little about these results even if they don't try to completely understand its proof. For example, Hamilton's 1982 result mentioned by PVAL is something I want to read about. Thanks for many good answers. Again, I emphasize that I have a very thin background in these things.</p>
[ { "answer_id": 258825, "score": 17, "text": "<p>In 1982, Richard Hamilton proved that a smooth closed 3-manifold admitting a metric with strictly positive Ricci curvature is a spherical space form (in particular such a manifold which is simply connected is $S^3$). This led to a (by all accounts still active) research program culminating in Grigori Perelman's proof of Thurston's geometrization conjecture in 2006 (and hence the Poincare conjecture). </p>\n" }, { "answer_id": 258833, "score": 12, "text": "<p>Donaldson's theorem uses the theory of instantons from mathematical physics to impose algebraic constraints on the possible intersection forms that a smooth 4-manifold can have (roughly speaking, they have to be diagonalizable over the integers). But Freedman had shown that there are a great many topological 4-manifolds whose intersection forms do not satisfy Donaldson's constraints, and consequently there are a lot of topological 4-manifolds which do not admit smooth structures.</p>\n\n<p>This led to vigorous ongoing activity in low dimensional topology. Seiberg-Witten theory simplified Donaldson's argument by replacing instantons with monopoles, and this theory gives most of the state-of-the art results in the area of exotic 4-manifolds. Floer used Donaldson's instantons to build a homology theory associated to a 3-manifold, and Donaldson noticed that a cobordism between 3-manifolds induces a map between their instanton homology groups, ultimately leading Atiyah and others to write down the axioms for what we now know as TQFT. Floer's idea was adapted to construct other homology theories for 3-manifolds, contact manifolds, symplectic manifolds, etc. and these tools are still proving state-of-the-art theorems in their areas.</p>\n" }, { "answer_id": 258822, "score": 9, "text": "<p>Let $G_1 \\subset G_2 \\subset \\dots$ be a sequence of groups and inclusion homomorphisms. It satisfies <em>homological stability</em> if for every $r \\in \\mathbb{N}_0$ there is an $n(r)$, such that for all $n &gt; n(r)$ the homomorphism \n$$\nH_r(G_n) \\to H_r(G_{n+1})\n$$\ninduced by the inclusion is an isomorphism. Examples of groups that have homological stability are the symmetric groups $S_n$, the braid groups $B_n$ and the automorphism group of the free groups $Aut(F_n)$. </p>\n\n<p>Harer proved that the mapping class groups $\\Gamma_{g,b}$ of surfaces $\\Sigma_{g,b}$ of genus $g$ with $b$ boundary components given by\n$$\n\\Gamma_{g,b} = \\pi_0(Diff^+(\\Sigma_{g,b}, \\partial \\Sigma_{g,b}))\n$$ \nsatisfy homological stability in the sense that the homomorphisms of the homology groups obtained from gluing on various surfaces along the boundary components eventually become isomorphisms. For example, gluing on a pair of pants for $b \\geq 1$ gives a homomorphism \n$$\n\\Gamma_{g,b} \\to \\Gamma_{g,b+1}\n$$\nwhich induces an isomorphism in homology $H_r(\\Gamma_{g,b}) \\to H_r(\\Gamma_{g,b+1})$ in the range $r &gt; 1$, $g \\geq 3r-2$. </p>\n\n<p>These results were later generalised by Madsen, Weiss, Galatius, Randal-Williams, Hatcher, Vogtmann, Wahl and many others. It has influenced the way to think about cobordism categories and has revealed a lot about the structure of moduli spaces of surfaces (possibly with tangential structures).</p>\n\n<p>This answer is much too short to give a huge and fruitful subject the attention it deserves. But maybe some of the people working in this area (some of which are active on mathoverflow) can expand on it.</p>\n" }, { "answer_id": 258834, "score": 8, "text": "<p>Haynes Miller's resolution of the Sullivan conjecture. </p>\n\n<p>Adams' solution if the Hopf invariant one problem. </p>\n" }, { "answer_id": 258837, "score": 4, "text": "<p><a href=\"https://en.wikipedia.org/wiki/Almgren%E2%80%93Pitts_min-max_theory\" rel=\"nofollow noreferrer\">Almgren–Pitts min-max theory</a>. It leads to the <a href=\"http://annals.math.princeton.edu/2014/179-2/p06\" rel=\"nofollow noreferrer\">proof of Willmore conjecture</a> by Marques and Neves.</p>\n" }, { "answer_id": 258841, "score": 4, "text": "<p>Following up the answer by Paul Siegel, in addition to Donaldson's spectacular result described by Paul, there are the papers of Sacks-Uhlenbeck on harmonic immersions of $2$-spheres into Riemannian manifolds and subsequent papers by Uhlenbeck on the \"bubbling\" phenomena for self-dual Yang-Mills connections and Taubes on gluing \"bubbles\" onto a self-dual Yang-Mills. These are the key technical results used by Donaldson in his thesis, as cited by Paul. They themselves laid the groundwork for a tremendous amount of work in geometric analysis since then.</p>\n\n<p>Here are only a few that I recall offhand:</p>\n\n<p>1) Minimal hypersurfaces (Schoen-Simon-Yau, Anderson)</p>\n\n<p>2) Einstein manifolds (Gao, Anderson-Cheeger)</p>\n\n<p>3) Recent work of Naber and Valtorta on stationary Yang-Mills connections</p>\n\n<p>4) Recent work of Sung Jin Oh showing that similar phenomenon exists for the hyperbolic Yang-Mills equations</p>\n" }, { "answer_id": 258846, "score": 4, "text": "<p>Hitchin's Self-duality paper.</p>\n\n<p>In this paper, Hitchin constructs a hyper-kähler structure on the moduli space of flat irreducible connections by identifying them with Higgs pairs (holomorphic bundle together with a holomorphic endomorphism valued 1-form).</p>\n\n<p>Initially, the construction was done for compact Riemann surfaces and bundles of rank 2, but generalize to Kähler manifolds, to punctured surfaces (parabolic structures) and to higher rank respectively pricipal bundles.</p>\n\n<p>The paper is not about proving one single theorem (he also introduces the integrable system nowadays know as the Hitchin system), I hope you can count this as an answer, as the subject of Higgs fields is still developing with many interesting connections to other subjects like mathematical physics or geometric langlands.</p>\n" }, { "answer_id": 258856, "score": 2, "text": "<p>Work done by many in low-dimensional topology that laid the groundwork for Thurston's work and his conjecture, which was proved by Perelman. I don't want to list names, since I'm sure I'll miss some key people.</p>\n" }, { "answer_id": 258825, "score": 17, "text": "<p>In 1982, Richard Hamilton proved that a smooth closed 3-manifold admitting a metric with strictly positive Ricci curvature is a spherical space form (in particular such a manifold which is simply connected is $S^3$). This led to a (by all accounts still active) research program culminating in Grigori Perelman's proof of Thurston's geometrization conjecture in 2006 (and hence the Poincare conjecture). </p>\n" }, { "answer_id": 258833, "score": 12, "text": "<p>Donaldson's theorem uses the theory of instantons from mathematical physics to impose algebraic constraints on the possible intersection forms that a smooth 4-manifold can have (roughly speaking, they have to be diagonalizable over the integers). But Freedman had shown that there are a great many topological 4-manifolds whose intersection forms do not satisfy Donaldson's constraints, and consequently there are a lot of topological 4-manifolds which do not admit smooth structures.</p>\n\n<p>This led to vigorous ongoing activity in low dimensional topology. Seiberg-Witten theory simplified Donaldson's argument by replacing instantons with monopoles, and this theory gives most of the state-of-the art results in the area of exotic 4-manifolds. Floer used Donaldson's instantons to build a homology theory associated to a 3-manifold, and Donaldson noticed that a cobordism between 3-manifolds induces a map between their instanton homology groups, ultimately leading Atiyah and others to write down the axioms for what we now know as TQFT. Floer's idea was adapted to construct other homology theories for 3-manifolds, contact manifolds, symplectic manifolds, etc. and these tools are still proving state-of-the-art theorems in their areas.</p>\n" }, { "answer_id": 258822, "score": 9, "text": "<p>Let $G_1 \\subset G_2 \\subset \\dots$ be a sequence of groups and inclusion homomorphisms. It satisfies <em>homological stability</em> if for every $r \\in \\mathbb{N}_0$ there is an $n(r)$, such that for all $n &gt; n(r)$ the homomorphism \n$$\nH_r(G_n) \\to H_r(G_{n+1})\n$$\ninduced by the inclusion is an isomorphism. Examples of groups that have homological stability are the symmetric groups $S_n$, the braid groups $B_n$ and the automorphism group of the free groups $Aut(F_n)$. </p>\n\n<p>Harer proved that the mapping class groups $\\Gamma_{g,b}$ of surfaces $\\Sigma_{g,b}$ of genus $g$ with $b$ boundary components given by\n$$\n\\Gamma_{g,b} = \\pi_0(Diff^+(\\Sigma_{g,b}, \\partial \\Sigma_{g,b}))\n$$ \nsatisfy homological stability in the sense that the homomorphisms of the homology groups obtained from gluing on various surfaces along the boundary components eventually become isomorphisms. For example, gluing on a pair of pants for $b \\geq 1$ gives a homomorphism \n$$\n\\Gamma_{g,b} \\to \\Gamma_{g,b+1}\n$$\nwhich induces an isomorphism in homology $H_r(\\Gamma_{g,b}) \\to H_r(\\Gamma_{g,b+1})$ in the range $r &gt; 1$, $g \\geq 3r-2$. </p>\n\n<p>These results were later generalised by Madsen, Weiss, Galatius, Randal-Williams, Hatcher, Vogtmann, Wahl and many others. It has influenced the way to think about cobordism categories and has revealed a lot about the structure of moduli spaces of surfaces (possibly with tangential structures).</p>\n\n<p>This answer is much too short to give a huge and fruitful subject the attention it deserves. But maybe some of the people working in this area (some of which are active on mathoverflow) can expand on it.</p>\n" }, { "answer_id": 258834, "score": 8, "text": "<p>Haynes Miller's resolution of the Sullivan conjecture. </p>\n\n<p>Adams' solution if the Hopf invariant one problem. </p>\n" }, { "answer_id": 258837, "score": 4, "text": "<p><a href=\"https://en.wikipedia.org/wiki/Almgren%E2%80%93Pitts_min-max_theory\" rel=\"nofollow noreferrer\">Almgren–Pitts min-max theory</a>. It leads to the <a href=\"http://annals.math.princeton.edu/2014/179-2/p06\" rel=\"nofollow noreferrer\">proof of Willmore conjecture</a> by Marques and Neves.</p>\n" }, { "answer_id": 258841, "score": 4, "text": "<p>Following up the answer by Paul Siegel, in addition to Donaldson's spectacular result described by Paul, there are the papers of Sacks-Uhlenbeck on harmonic immersions of $2$-spheres into Riemannian manifolds and subsequent papers by Uhlenbeck on the \"bubbling\" phenomena for self-dual Yang-Mills connections and Taubes on gluing \"bubbles\" onto a self-dual Yang-Mills. These are the key technical results used by Donaldson in his thesis, as cited by Paul. They themselves laid the groundwork for a tremendous amount of work in geometric analysis since then.</p>\n\n<p>Here are only a few that I recall offhand:</p>\n\n<p>1) Minimal hypersurfaces (Schoen-Simon-Yau, Anderson)</p>\n\n<p>2) Einstein manifolds (Gao, Anderson-Cheeger)</p>\n\n<p>3) Recent work of Naber and Valtorta on stationary Yang-Mills connections</p>\n\n<p>4) Recent work of Sung Jin Oh showing that similar phenomenon exists for the hyperbolic Yang-Mills equations</p>\n" }, { "answer_id": 258846, "score": 4, "text": "<p>Hitchin's Self-duality paper.</p>\n\n<p>In this paper, Hitchin constructs a hyper-kähler structure on the moduli space of flat irreducible connections by identifying them with Higgs pairs (holomorphic bundle together with a holomorphic endomorphism valued 1-form).</p>\n\n<p>Initially, the construction was done for compact Riemann surfaces and bundles of rank 2, but generalize to Kähler manifolds, to punctured surfaces (parabolic structures) and to higher rank respectively pricipal bundles.</p>\n\n<p>The paper is not about proving one single theorem (he also introduces the integrable system nowadays know as the Hitchin system), I hope you can count this as an answer, as the subject of Higgs fields is still developing with many interesting connections to other subjects like mathematical physics or geometric langlands.</p>\n" }, { "answer_id": 258856, "score": 2, "text": "<p>Work done by many in low-dimensional topology that laid the groundwork for Thurston's work and his conjecture, which was proved by Perelman. I don't want to list names, since I'm sure I'll miss some key people.</p>\n" }, { "answer_id": 258825, "score": 17, "text": "<p>In 1982, Richard Hamilton proved that a smooth closed 3-manifold admitting a metric with strictly positive Ricci curvature is a spherical space form (in particular such a manifold which is simply connected is $S^3$). This led to a (by all accounts still active) research program culminating in Grigori Perelman's proof of Thurston's geometrization conjecture in 2006 (and hence the Poincare conjecture). </p>\n" }, { "answer_id": 258833, "score": 12, "text": "<p>Donaldson's theorem uses the theory of instantons from mathematical physics to impose algebraic constraints on the possible intersection forms that a smooth 4-manifold can have (roughly speaking, they have to be diagonalizable over the integers). But Freedman had shown that there are a great many topological 4-manifolds whose intersection forms do not satisfy Donaldson's constraints, and consequently there are a lot of topological 4-manifolds which do not admit smooth structures.</p>\n\n<p>This led to vigorous ongoing activity in low dimensional topology. Seiberg-Witten theory simplified Donaldson's argument by replacing instantons with monopoles, and this theory gives most of the state-of-the art results in the area of exotic 4-manifolds. Floer used Donaldson's instantons to build a homology theory associated to a 3-manifold, and Donaldson noticed that a cobordism between 3-manifolds induces a map between their instanton homology groups, ultimately leading Atiyah and others to write down the axioms for what we now know as TQFT. Floer's idea was adapted to construct other homology theories for 3-manifolds, contact manifolds, symplectic manifolds, etc. and these tools are still proving state-of-the-art theorems in their areas.</p>\n" }, { "answer_id": 258822, "score": 9, "text": "<p>Let $G_1 \\subset G_2 \\subset \\dots$ be a sequence of groups and inclusion homomorphisms. It satisfies <em>homological stability</em> if for every $r \\in \\mathbb{N}_0$ there is an $n(r)$, such that for all $n &gt; n(r)$ the homomorphism \n$$\nH_r(G_n) \\to H_r(G_{n+1})\n$$\ninduced by the inclusion is an isomorphism. Examples of groups that have homological stability are the symmetric groups $S_n$, the braid groups $B_n$ and the automorphism group of the free groups $Aut(F_n)$. </p>\n\n<p>Harer proved that the mapping class groups $\\Gamma_{g,b}$ of surfaces $\\Sigma_{g,b}$ of genus $g$ with $b$ boundary components given by\n$$\n\\Gamma_{g,b} = \\pi_0(Diff^+(\\Sigma_{g,b}, \\partial \\Sigma_{g,b}))\n$$ \nsatisfy homological stability in the sense that the homomorphisms of the homology groups obtained from gluing on various surfaces along the boundary components eventually become isomorphisms. For example, gluing on a pair of pants for $b \\geq 1$ gives a homomorphism \n$$\n\\Gamma_{g,b} \\to \\Gamma_{g,b+1}\n$$\nwhich induces an isomorphism in homology $H_r(\\Gamma_{g,b}) \\to H_r(\\Gamma_{g,b+1})$ in the range $r &gt; 1$, $g \\geq 3r-2$. </p>\n\n<p>These results were later generalised by Madsen, Weiss, Galatius, Randal-Williams, Hatcher, Vogtmann, Wahl and many others. It has influenced the way to think about cobordism categories and has revealed a lot about the structure of moduli spaces of surfaces (possibly with tangential structures).</p>\n\n<p>This answer is much too short to give a huge and fruitful subject the attention it deserves. But maybe some of the people working in this area (some of which are active on mathoverflow) can expand on it.</p>\n" }, { "answer_id": 258834, "score": 8, "text": "<p>Haynes Miller's resolution of the Sullivan conjecture. </p>\n\n<p>Adams' solution if the Hopf invariant one problem. </p>\n" }, { "answer_id": 258837, "score": 4, "text": "<p><a href=\"https://en.wikipedia.org/wiki/Almgren%E2%80%93Pitts_min-max_theory\" rel=\"nofollow noreferrer\">Almgren–Pitts min-max theory</a>. It leads to the <a href=\"http://annals.math.princeton.edu/2014/179-2/p06\" rel=\"nofollow noreferrer\">proof of Willmore conjecture</a> by Marques and Neves.</p>\n" }, { "answer_id": 258841, "score": 4, "text": "<p>Following up the answer by Paul Siegel, in addition to Donaldson's spectacular result described by Paul, there are the papers of Sacks-Uhlenbeck on harmonic immersions of $2$-spheres into Riemannian manifolds and subsequent papers by Uhlenbeck on the \"bubbling\" phenomena for self-dual Yang-Mills connections and Taubes on gluing \"bubbles\" onto a self-dual Yang-Mills. These are the key technical results used by Donaldson in his thesis, as cited by Paul. They themselves laid the groundwork for a tremendous amount of work in geometric analysis since then.</p>\n\n<p>Here are only a few that I recall offhand:</p>\n\n<p>1) Minimal hypersurfaces (Schoen-Simon-Yau, Anderson)</p>\n\n<p>2) Einstein manifolds (Gao, Anderson-Cheeger)</p>\n\n<p>3) Recent work of Naber and Valtorta on stationary Yang-Mills connections</p>\n\n<p>4) Recent work of Sung Jin Oh showing that similar phenomenon exists for the hyperbolic Yang-Mills equations</p>\n" }, { "answer_id": 258846, "score": 4, "text": "<p>Hitchin's Self-duality paper.</p>\n\n<p>In this paper, Hitchin constructs a hyper-kähler structure on the moduli space of flat irreducible connections by identifying them with Higgs pairs (holomorphic bundle together with a holomorphic endomorphism valued 1-form).</p>\n\n<p>Initially, the construction was done for compact Riemann surfaces and bundles of rank 2, but generalize to Kähler manifolds, to punctured surfaces (parabolic structures) and to higher rank respectively pricipal bundles.</p>\n\n<p>The paper is not about proving one single theorem (he also introduces the integrable system nowadays know as the Hitchin system), I hope you can count this as an answer, as the subject of Higgs fields is still developing with many interesting connections to other subjects like mathematical physics or geometric langlands.</p>\n" }, { "answer_id": 258856, "score": 2, "text": "<p>Work done by many in low-dimensional topology that laid the groundwork for Thurston's work and his conjecture, which was proved by Perelman. I don't want to list names, since I'm sure I'll miss some key people.</p>\n" } ]
Question: Theorems that led to very successful research programs in Geometry and Topology Answer 1: <p>In 1982, Richard Hamilton proved that a smooth closed 3-manifold admitting a metric with strictly positive Ricci curvature is a spherical space form (in particular such a manifold which is simply connected is $S^3$). This led to a (by all accounts still active) research program culminating in Grigori Perelman's proof of Thurston's geometrization conjecture in 2006 (and hence the Poincare conjecture). </p> Answer 2: <p>Donaldson's theorem uses the theory of instantons from mathematical physics to impose algebraic constraints on the possible intersection forms that a smooth 4-manifold can have (roughly speaking, they have to be diagonalizable over the integers). But Freedman had shown that there are a great many topological 4-manifolds whose intersection forms do not satisfy Donaldson's constraints, and consequently there are a lot of topological 4-manifolds which do not admit smooth structures.</p> <p>This led to vigorous ongoing activity in low dimensional topology. Seiberg-Witten theory simplified Donaldson's argument by replacing instantons with monopoles, and this theory gives most of the state-of-the art results in the area of exotic 4-manifolds. Floer used Donaldson's instantons to build a homology theory associated to a 3-manifold, and Donaldson noticed that a cobordism between 3-manifolds induces a map between their instanton homology groups, ultimately leading Atiyah and others to write down the axioms for what we now know as TQFT. Floer's idea was adapted to construct other homology theories for 3-manifolds, contact manifolds, symplectic manifolds, etc. and these tools are still proving state-of-the-art theorems in their areas.</p> Answer 3: <p>Let $G_1 \subset G_2 \subset \dots$ be a sequence of groups and inclusion homomorphisms. It satisfies <em>homological stability</em> if for every $r \in \mathbb{N}_0$ there is an $n(r)$, such that for all $n &gt; n(r)$ the homomorphism $$ H_r(G_n) \to H_r(G_{n+1}) $$ induced by the inclusion is an isomorphism. Examples of groups that have homological stability are the symmetric groups $S_n$, the braid groups $B_n$ and the automorphism group of the free groups $Aut(F_n)$. </p> <p>Harer proved that the mapping class groups $\Gamma_{g,b}$ of surfaces $\Sigma_{g,b}$ of genus $g$ with $b$ boundary components given by $$ \Gamma_{g,b} = \pi_0(Diff^+(\Sigma_{g,b}, \partial \Sigma_{g,b})) $$ satisfy homological stability in the sense that the homomorphisms of the homology groups obtained from gluing on various surfaces along the boundary components eventually become isomorphisms. For example, gluing on a pair of pants for $b \geq 1$ gives a homomorphism $$ \Gamma_{g,b} \to \Gamma_{g,b+1} $$ which induces an isomorphism in homology $H_r(\Gamma_{g,b}) \to H_r(\Gamma_{g,b+1})$ in the range $r &gt; 1$, $g \geq 3r-2$. </p> <p>These results were later generalised by Madsen, Weiss, Galatius, Randal-Williams, Hatcher, Vogtmann, Wahl and many others. It has influenced the way to think about cobordism categories and has revealed a lot about the structure of moduli spaces of surfaces (possibly with tangential structures).</p> <p>This answer is much too short to give a huge and fruitful subject the attention it deserves. But maybe some of the people working in this area (some of which are active on mathoverflow) can expand on it.</p> Answer 4: <p>Haynes Miller's resolution of the Sullivan conjecture. </p> <p>Adams' solution if the Hopf invariant one problem. </p> Answer 5: <p><a href="https://en.wikipedia.org/wiki/Almgren%E2%80%93Pitts_min-max_theory" rel="nofollow noreferrer">Almgren–Pitts min-max theory</a>. It leads to the <a href="http://annals.math.princeton.edu/2014/179-2/p06" rel="nofollow noreferrer">proof of Willmore conjecture</a> by Marques and Neves.</p> Answer 6: <p>Following up the answer by Paul Siegel, in addition to Donaldson's spectacular result described by Paul, there are the papers of Sacks-Uhlenbeck on harmonic immersions of $2$-spheres into Riemannian manifolds and subsequent papers by Uhlenbeck on the "bubbling" phenomena for self-dual Yang-Mills connections and Taubes on gluing "bubbles" onto a self-dual Yang-Mills. These are the key technical results used by Donaldson in his thesis, as cited by Paul. They themselves laid the groundwork for a tremendous amount of work in geometric analysis since then.</p> <p>Here are only a few that I recall offhand:</p> <p>1) Minimal hypersurfaces (Schoen-Simon-Yau, Anderson)</p> <p>2) Einstein manifolds (Gao, Anderson-Cheeger)</p> <p>3) Recent work of Naber and Valtorta on stationary Yang-Mills connections</p> <p>4) Recent work of Sung Jin Oh showing that similar phenomenon exists for the hyperbolic Yang-Mills equations</p> Answer 7: <p>Hitchin's Self-duality paper.</p> <p>In this paper, Hitchin constructs a hyper-kähler structure on the moduli space of flat irreducible connections by identifying them with Higgs pairs (holomorphic bundle together with a holomorphic endomorphism valued 1-form).</p> <p>Initially, the construction was done for compact Riemann surfaces and bundles of rank 2, but generalize to Kähler manifolds, to punctured surfaces (parabolic structures) and to higher rank respectively pricipal bundles.</p> <p>The paper is not about proving one single theorem (he also introduces the integrable system nowadays know as the Hitchin system), I hope you can count this as an answer, as the subject of Higgs fields is still developing with many interesting connections to other subjects like mathematical physics or geometric langlands.</p> Answer 8: <p>Work done by many in low-dimensional topology that laid the groundwork for Thurston's work and his conjecture, which was proved by Perelman. I don't want to list names, since I'm sure I'll miss some key people.</p> Answer 9: <p>In 1982, Richard Hamilton proved that a smooth closed 3-manifold admitting a metric with strictly positive Ricci curvature is a spherical space form (in particular such a manifold which is simply connected is $S^3$). This led to a (by all accounts still active) research program culminating in Grigori Perelman's proof of Thurston's geometrization conjecture in 2006 (and hence the Poincare conjecture). </p> Answer 10: <p>Donaldson's theorem uses the theory of instantons from mathematical physics to impose algebraic constraints on the possible intersection forms that a smooth 4-manifold can have (roughly speaking, they have to be diagonalizable over the integers). But Freedman had shown that there are a great many topological 4-manifolds whose intersection forms do not satisfy Donaldson's constraints, and consequently there are a lot of topological 4-manifolds which do not admit smooth structures.</p> <p>This led to vigorous ongoing activity in low dimensional topology. Seiberg-Witten theory simplified Donaldson's argument by replacing instantons with monopoles, and this theory gives most of the state-of-the art results in the area of exotic 4-manifolds. Floer used Donaldson's instantons to build a homology theory associated to a 3-manifold, and Donaldson noticed that a cobordism between 3-manifolds induces a map between their instanton homology groups, ultimately leading Atiyah and others to write down the axioms for what we now know as TQFT. Floer's idea was adapted to construct other homology theories for 3-manifolds, contact manifolds, symplectic manifolds, etc. and these tools are still proving state-of-the-art theorems in their areas.</p> Answer 11: <p>Let $G_1 \subset G_2 \subset \dots$ be a sequence of groups and inclusion homomorphisms. It satisfies <em>homological stability</em> if for every $r \in \mathbb{N}_0$ there is an $n(r)$, such that for all $n &gt; n(r)$ the homomorphism $$ H_r(G_n) \to H_r(G_{n+1}) $$ induced by the inclusion is an isomorphism. Examples of groups that have homological stability are the symmetric groups $S_n$, the braid groups $B_n$ and the automorphism group of the free groups $Aut(F_n)$. </p> <p>Harer proved that the mapping class groups $\Gamma_{g,b}$ of surfaces $\Sigma_{g,b}$ of genus $g$ with $b$ boundary components given by $$ \Gamma_{g,b} = \pi_0(Diff^+(\Sigma_{g,b}, \partial \Sigma_{g,b})) $$ satisfy homological stability in the sense that the homomorphisms of the homology groups obtained from gluing on various surfaces along the boundary components eventually become isomorphisms. For example, gluing on a pair of pants for $b \geq 1$ gives a homomorphism $$ \Gamma_{g,b} \to \Gamma_{g,b+1} $$ which induces an isomorphism in homology $H_r(\Gamma_{g,b}) \to H_r(\Gamma_{g,b+1})$ in the range $r &gt; 1$, $g \geq 3r-2$. </p> <p>These results were later generalised by Madsen, Weiss, Galatius, Randal-Williams, Hatcher, Vogtmann, Wahl and many others. It has influenced the way to think about cobordism categories and has revealed a lot about the structure of moduli spaces of surfaces (possibly with tangential structures).</p> <p>This answer is much too short to give a huge and fruitful subject the attention it deserves. But maybe some of the people working in this area (some of which are active on mathoverflow) can expand on it.</p> Answer 12: <p>Haynes Miller's resolution of the Sullivan conjecture. </p> <p>Adams' solution if the Hopf invariant one problem. </p> Answer 13: <p><a href="https://en.wikipedia.org/wiki/Almgren%E2%80%93Pitts_min-max_theory" rel="nofollow noreferrer">Almgren–Pitts min-max theory</a>. It leads to the <a href="http://annals.math.princeton.edu/2014/179-2/p06" rel="nofollow noreferrer">proof of Willmore conjecture</a> by Marques and Neves.</p> Answer 14: <p>Following up the answer by Paul Siegel, in addition to Donaldson's spectacular result described by Paul, there are the papers of Sacks-Uhlenbeck on harmonic immersions of $2$-spheres into Riemannian manifolds and subsequent papers by Uhlenbeck on the "bubbling" phenomena for self-dual Yang-Mills connections and Taubes on gluing "bubbles" onto a self-dual Yang-Mills. These are the key technical results used by Donaldson in his thesis, as cited by Paul. They themselves laid the groundwork for a tremendous amount of work in geometric analysis since then.</p> <p>Here are only a few that I recall offhand:</p> <p>1) Minimal hypersurfaces (Schoen-Simon-Yau, Anderson)</p> <p>2) Einstein manifolds (Gao, Anderson-Cheeger)</p> <p>3) Recent work of Naber and Valtorta on stationary Yang-Mills connections</p> <p>4) Recent work of Sung Jin Oh showing that similar phenomenon exists for the hyperbolic Yang-Mills equations</p> Answer 15: <p>Hitchin's Self-duality paper.</p> <p>In this paper, Hitchin constructs a hyper-kähler structure on the moduli space of flat irreducible connections by identifying them with Higgs pairs (holomorphic bundle together with a holomorphic endomorphism valued 1-form).</p> <p>Initially, the construction was done for compact Riemann surfaces and bundles of rank 2, but generalize to Kähler manifolds, to punctured surfaces (parabolic structures) and to higher rank respectively pricipal bundles.</p> <p>The paper is not about proving one single theorem (he also introduces the integrable system nowadays know as the Hitchin system), I hope you can count this as an answer, as the subject of Higgs fields is still developing with many interesting connections to other subjects like mathematical physics or geometric langlands.</p> Answer 16: <p>Work done by many in low-dimensional topology that laid the groundwork for Thurston's work and his conjecture, which was proved by Perelman. I don't want to list names, since I'm sure I'll miss some key people.</p> Answer 17: <p>In 1982, Richard Hamilton proved that a smooth closed 3-manifold admitting a metric with strictly positive Ricci curvature is a spherical space form (in particular such a manifold which is simply connected is $S^3$). This led to a (by all accounts still active) research program culminating in Grigori Perelman's proof of Thurston's geometrization conjecture in 2006 (and hence the Poincare conjecture). </p> Answer 18: <p>Donaldson's theorem uses the theory of instantons from mathematical physics to impose algebraic constraints on the possible intersection forms that a smooth 4-manifold can have (roughly speaking, they have to be diagonalizable over the integers). But Freedman had shown that there are a great many topological 4-manifolds whose intersection forms do not satisfy Donaldson's constraints, and consequently there are a lot of topological 4-manifolds which do not admit smooth structures.</p> <p>This led to vigorous ongoing activity in low dimensional topology. Seiberg-Witten theory simplified Donaldson's argument by replacing instantons with monopoles, and this theory gives most of the state-of-the art results in the area of exotic 4-manifolds. Floer used Donaldson's instantons to build a homology theory associated to a 3-manifold, and Donaldson noticed that a cobordism between 3-manifolds induces a map between their instanton homology groups, ultimately leading Atiyah and others to write down the axioms for what we now know as TQFT. Floer's idea was adapted to construct other homology theories for 3-manifolds, contact manifolds, symplectic manifolds, etc. and these tools are still proving state-of-the-art theorems in their areas.</p> Answer 19: <p>Let $G_1 \subset G_2 \subset \dots$ be a sequence of groups and inclusion homomorphisms. It satisfies <em>homological stability</em> if for every $r \in \mathbb{N}_0$ there is an $n(r)$, such that for all $n &gt; n(r)$ the homomorphism $$ H_r(G_n) \to H_r(G_{n+1}) $$ induced by the inclusion is an isomorphism. Examples of groups that have homological stability are the symmetric groups $S_n$, the braid groups $B_n$ and the automorphism group of the free groups $Aut(F_n)$. </p> <p>Harer proved that the mapping class groups $\Gamma_{g,b}$ of surfaces $\Sigma_{g,b}$ of genus $g$ with $b$ boundary components given by $$ \Gamma_{g,b} = \pi_0(Diff^+(\Sigma_{g,b}, \partial \Sigma_{g,b})) $$ satisfy homological stability in the sense that the homomorphisms of the homology groups obtained from gluing on various surfaces along the boundary components eventually become isomorphisms. For example, gluing on a pair of pants for $b \geq 1$ gives a homomorphism $$ \Gamma_{g,b} \to \Gamma_{g,b+1} $$ which induces an isomorphism in homology $H_r(\Gamma_{g,b}) \to H_r(\Gamma_{g,b+1})$ in the range $r &gt; 1$, $g \geq 3r-2$. </p> <p>These results were later generalised by Madsen, Weiss, Galatius, Randal-Williams, Hatcher, Vogtmann, Wahl and many others. It has influenced the way to think about cobordism categories and has revealed a lot about the structure of moduli spaces of surfaces (possibly with tangential structures).</p> <p>This answer is much too short to give a huge and fruitful subject the attention it deserves. But maybe some of the people working in this area (some of which are active on mathoverflow) can expand on it.</p> Answer 20: <p>Haynes Miller's resolution of the Sullivan conjecture. </p> <p>Adams' solution if the Hopf invariant one problem. </p> Answer 21: <p><a href="https://en.wikipedia.org/wiki/Almgren%E2%80%93Pitts_min-max_theory" rel="nofollow noreferrer">Almgren–Pitts min-max theory</a>. It leads to the <a href="http://annals.math.princeton.edu/2014/179-2/p06" rel="nofollow noreferrer">proof of Willmore conjecture</a> by Marques and Neves.</p> Answer 22: <p>Following up the answer by Paul Siegel, in addition to Donaldson's spectacular result described by Paul, there are the papers of Sacks-Uhlenbeck on harmonic immersions of $2$-spheres into Riemannian manifolds and subsequent papers by Uhlenbeck on the "bubbling" phenomena for self-dual Yang-Mills connections and Taubes on gluing "bubbles" onto a self-dual Yang-Mills. These are the key technical results used by Donaldson in his thesis, as cited by Paul. They themselves laid the groundwork for a tremendous amount of work in geometric analysis since then.</p> <p>Here are only a few that I recall offhand:</p> <p>1) Minimal hypersurfaces (Schoen-Simon-Yau, Anderson)</p> <p>2) Einstein manifolds (Gao, Anderson-Cheeger)</p> <p>3) Recent work of Naber and Valtorta on stationary Yang-Mills connections</p> <p>4) Recent work of Sung Jin Oh showing that similar phenomenon exists for the hyperbolic Yang-Mills equations</p> Answer 23: <p>Hitchin's Self-duality paper.</p> <p>In this paper, Hitchin constructs a hyper-kähler structure on the moduli space of flat irreducible connections by identifying them with Higgs pairs (holomorphic bundle together with a holomorphic endomorphism valued 1-form).</p> <p>Initially, the construction was done for compact Riemann surfaces and bundles of rank 2, but generalize to Kähler manifolds, to punctured surfaces (parabolic structures) and to higher rank respectively pricipal bundles.</p> <p>The paper is not about proving one single theorem (he also introduces the integrable system nowadays know as the Hitchin system), I hope you can count this as an answer, as the subject of Higgs fields is still developing with many interesting connections to other subjects like mathematical physics or geometric langlands.</p> Answer 24: <p>Work done by many in low-dimensional topology that laid the groundwork for Thurston's work and his conjecture, which was proved by Perelman. I don't want to list names, since I'm sure I'll miss some key people.</p>
Explicit invariant of tensors nonvanishing on the diagonal
259,106
18
https://mathoverflow.net/questions/259106/explicit-invariant-of-tensors-nonvanishing-on-the-diagonal
<p>The group $SL_n \times SL_n \times SL_n$ acts naturally on the vector space $\mathbb C^n \otimes \mathbb C^n \otimes \mathbb C^n$ and has a rather large ring of polynomial invariants. The element $$\sum_{i=1}^n e_i \otimes e_i \otimes e_i \in \mathbb C^n \otimes \mathbb C^n \otimes \mathbb C^n$$ is known to be GIT-semistable with respect to this action. In other words, there is a homogeneous $SL_n \times SL_n \times SL_n$-invariant polynomial, of nonzero degree, which is nonvanishing on this element. The proof I know (<a href="https://arxiv.org/pdf/1605.06702v3.pdf" rel="noreferrer">Theorem 4.7</a>) uses the Hilbert-Mumford criterion and so does not explicitly construct the polynomial.</p> <blockquote> <p>Can you give an explicit homogeneous $SL_n \times SL_n \times SL_n$-invariant polynomial that is nonzero on this element.</p> </blockquote> <p>One can check from the definition of the hyperdeterminant that the hyperdeterminant vanishes on this element as long as $n&gt;2$. So that won't work.</p> <p>I don't know how many other natural ways there are for defining an invariant function on $\mathbb C^n \otimes \mathbb C^n \otimes \mathbb C^n$ for all $n$ simultaneously there are.</p> <p>My motivation for this problem is to understand semistability of tensor powers of tensors better, and so to understand slice rank of tensor powers of tensors better. I want to give explicit criteria to show that all a tensor's tensor powers are semistable. Tensor powers of the diagonal tensor remain, so the diagonal tensor certainly has this property. Given an explicit invariant, perhaps one could find an explicit neighborhood of the diagonal tensor consisting of tensors that also have this property.</p>
[ { "answer_id": 259195, "score": 9, "text": "<p>Let me start with some remarks about the classical symbolic method (without which one cannot understand 19th century invariant theory)\nand multisymmetric functions.\nI will use an example first. Take four series of three variables $a=(a_1,a_2,a_3)$, $b=(b_1,b_2,b_3)$,\n$c=(c_1,c_2,c_3)$ and $d=(d_1,d_2,d_3)$.\nNow define the polynomial $\\mathcal{A}$ in these 12 indeterminates given by\n$$\n\\mathcal{A}=(bcd)(acd)(abd)(abc)\n$$\nwhere we used a shorthand notation for determinants\n$$\n(bcd)=\\left|\n\\begin{array}{ccc}\nb_1 &amp; b_2 &amp; b_3 \\\\\nc_1 &amp; c_2 &amp; c_3 \\\\\nd_1 &amp; d_2 &amp; d_3 \\\\\n\\end{array}\n\\right|\n$$\nand similarly for the other \"bracket factors\" $(acd)$, etc.\nThis is an example of multisymmetric function in the \"letters\" $a,b,c,d$ which is multihomogeneous of multidegree $[3,3,3,3]$.\nLet me write $M_{[3,3,3,3]}^3$ for the space of such multisymmetric functions. The subscript is the multidegree while the superscript refers\nto the number of variables in each series. \nThe symmetric group acting is $\\mathfrak{S}_4$ which permutes these series. If each series had one variable in it instead of\nthree we would be in the familiar realm of the theory of ordinary symmetric functions. The analogous theory of multisymmetric\nfunctions is much more complicated and has been studied classically by Euler, Lagrange, Poisson, Schl&auml;fli, Brill, Gordan, Junker\nand more recently by Angeniol, Dalbec, Briand, Vaccarino, Domokos and others.\nA nice yet old reference on this is the book on elimination theory by Faa di Bruno I mentioned in <a href=\"https://mathoverflow.net/questions/117415/old-books-still-used/117417#117417\">this MO answer</a>.</p>\n\n<p>Consider now a generic ternary cubic form $F=F(x_1,x_2,x_3)$. One can write it as\n$$\nF(x)=\\sum_{i_1,i_2,i_3=1}^3 F_{i_1,i_2,i_3} x_{i_1} x_{i_2} x_{i_3}\n$$\nwhere $(F_{i_1,i_2,i_3})$ is a symmetric tensor (by which I mean a \"matrix\" rather than a Bourbaki tensor).\nThis form lives in the symmetric power $S^3(\\mathbb{C}^3)$ but I will be (intentionally) sloppy and write $S^3$ instead.\nThe space of degree 4 homogeneous polynomials in the coefficients of $F$ will be denoted by $S^4(S^3)$.</p>\n\n<p>Let me define a linear map $\\Phi:S^4(S^3)\\rightarrow M_{[3,3,3,3]}^3$ as follows.\nGiven a degree four homogeneous polynomial $C(F)$ first construct its full polarization\n$$\n\\widetilde{C}(F_1,F_2,F_3,F_4)=\\frac{1}{4!} \\frac{\\partial^4}{\\partial t_1\\partial t_2\\partial t_3\\partial t_4}\nC(t_1F_1+t_2 F_2+t_3 F_3+t_4 F_4)\n$$ \nwhich satisfies $\\widetilde{C}(F,F,F,F)=C(F)$. \nNow set $F_1(x)=a_{x}^3$ where $a_x$ is the classical notation for the linear form $a_1x_1+a_2x_2+a_3x_3$ and similarly\nspecialize to the cubes of linear forms $F_2(x)=b_{x}^3$, $F_3(x)=c_{x}^3$ and $F_4(x)=d_{x}^3$.\nFinally, define $\\Phi(C)=\\widetilde{C}(F_1,F_2,F_3,F_4)$.</p>\n\n<p>I will also define another linear map $\\Psi: M_{[3,3,3,3]}^3\\rightarrow S^4(S^3)$.\nLet $\\mathcal{S}=\\mathcal{S}(a,b,c,d)$ be a multisymmetric function in $M_{[3,3,3,3]}^3$.\nWe define the differential operator\n$$\n\\mathcal{D}=\\frac{1}{3!^4} \nF\\left(\\frac{\\partial}{\\partial a_1},\\frac{\\partial}{\\partial a_2},\\frac{\\partial}{\\partial a_3}\\right)\nF\\left(\\frac{\\partial}{\\partial b_1},\\frac{\\partial}{\\partial b_2},\\frac{\\partial}{\\partial b_3}\\right)\n$$\n$$\nF\\left(\\frac{\\partial}{\\partial c_1},\\frac{\\partial}{\\partial c_2},\\frac{\\partial}{\\partial c_3}\\right)\nF\\left(\\frac{\\partial}{\\partial d_1},\\frac{\\partial}{\\partial d_2},\\frac{\\partial}{\\partial d_3}\\right)\n$$\nand let\n$\\Psi(\\mathcal{S})=\\mathcal{D}\\ \\mathcal{S}$, i.e., the result of applying the differential operator $\\mathcal{D}$\nto the polynomial\n$\\mathcal{S}$.</p>\n\n<p>It is not hard to see that $\\Phi$ and $\\Psi$ are inverses of one another.\nIn the particular case of $\\mathcal{A}$ above, the image $\\Psi(\\mathcal{A})$ is Aronhold's degree four $SL_3$ invariant of plane cubics\nand $\\mathcal{A}$ itself is called the (German) symbolic form of that invariant. The classics usually omitted \"$\\Psi(\\cdots)$\" but they \nknew perfectly well what they were doing. Also note that this formalism, in an essentially equivalent form, was invented earlier by Cayley\nas I briefly explained in Section 1.11.4 of my article\n<a href=\"https://arxiv.org/abs/1604.05259\" rel=\"nofollow noreferrer\">\"A Second-Quantized Kolmogorov-Chentsov Theorem\"</a>.\nOne can easily write $\\Psi(\\mathcal{A})$ applied to $F$, in \"physicsy\" tensor contraction notation as\n$$\n\\epsilon_{i_{1} i_{2} i_{3}}\\epsilon_{i_{4} i_{5} i_{6}}\\epsilon_{i_{7} i_{8} i_{9}}\\epsilon_{i_{10} i_{11} i_{12}}\nF_{i_{4} i_{7} i_{10}} F_{i_{1} i_{8} i_{11}} F_{i_{2} i_{5} i_{12}} F_{i_{3} i_{6} i_{9}}\n$$\nwith summation over all twelve indices in the set $\\{1,2,3\\}$. I used the standard physics notation for the Levi-Civita tensor\n$\\epsilon_{i_1 i_2 i_3}$ equal to zero if some indices are equal and otherwise to the sign of the permutation $i_1,i_2,i_3$\nof $1,2,3$. This notation is consistent with the one I used in my previous answers to these MO questions:\n<a href=\"https://mathoverflow.net/questions/121715/invariant-polynomials-for-a-product-of-algebraic-groups\">Q1</a>,\n<a href=\"https://mathoverflow.net/questions/257709/ring-of-invariants-of-operatornamesl-6-acting-on-lambda3-mathbb-c6\">Q2</a>,\nand <a href=\"https://mathoverflow.net/questions/255492/how-to-constructively-combinatorially-prove-schur-weyl-duality\">Q3</a>.\nAt bottom, this is an intrinsically diagrammatic language and if the reader had trouble following what I said so far, it's probably because\nI did not draw any pictures or \"finite-dimensional Feynman diagrams\" (I don't know how to do this on MO). To see these pictures,\nplease have a look at <a href=\"https://arxiv.org/abs/1402.2669\" rel=\"nofollow noreferrer\">this article</a> about an invariant of cubic threefolds or Section 2\nof <a href=\"https://arxiv.org/abs/0904.1734\" rel=\"nofollow noreferrer\">this one</a> mostly on binary forms.\nIn a nutshell, the above contraction can be encoded in a bipartite graph with 4 $\\epsilon$-vertices and 4 $F$-vertices, all of degree three.</p>\n\n<p>Of course the above formalism generalizes with no difficulty (except managing notations) to plethysms of symmetric powers\n$S^p(S^q(\\mathbb{C}^n))$ and similar $\\Phi$, $\\Psi$ maps giving the isomorphism with the relevant space of multisymmetric functions, i.e.,\n$M_{[q,q,\\ldots,q]}^{n}$ (with $q$ repeated $p$ times) or in simpler notation $M_{[q^p]}^{n}$.\nThere is a canonical $GL_n$-equivariant map\n$$\nH_{p,q}:S^p(S^q(\\mathbb{C}^n))\\rightarrow S^q(S^p(\\mathbb{C}^n))\n$$ \nwith \"H\" standing for Hermite, Hadamard or Howe. I will not define it here but it's the first thing that would come to mind\nif asked to produce such a map\n(see <a href=\"https://arxiv.org/abs/1305.7387\" rel=\"nofollow noreferrer\">this review</a> by Landsberg for a detailed discussion in the wider context of geometric complexity theory).\nIt is not hard to see that a homogeneous multisymmetric function in $M_{[p^q]}^{n}\\simeq S^q(S^p(\\mathbb{C}^n))$ is in the image\nof $H_{p,q}$ iff it is a polynomial in the homogeneous elementary multisymmetric functions which belong to $M_{[1^q]}^{n}$.\nIn the context of the Aronhold invariant above, these elementary functions are the coefficients of the polynomial\n$$\nP(x_1,x_2,x_3)=a_x b_x c_x d_x\n$$ \nand similarly in general.</p>\n\n<p>Finally, I can start discussing Will's excellent question. Let $T_{i_1,i_2,i_3}$ denote the tensor of interest with the action of\n$(g,h,k)\\in (SL_n)^3$ given by\n$$\n[(g,h,k)\\cdot T]_{i_1,i_2,i_3}=\\sum_{j_1,j_2,j_3=1}^{n} g_{i_1 j_1} h_{i_2 j_2} k_{i_3 j_3} T_{j_1,j_2,j_3}\\ .\n$$\nIf Will's tensor had an even number $r$ of indices with the action of $(SL_n)^r$ the answer would be trivial.\nTake $n$ copies of $T$, contract all first indices of these copies with one $\\epsilon$,\nthen all second indices with another $\\epsilon$, etc. For $r=3$ this tentative invariant vanishes identically.\nOne can however keep the first indices free, and this gives a map $T\\mapsto \\Gamma(T)=F$ producing an $n$-ary $n$-ic\n$F(x_1,\\ldots,x_n)$ corresponding to the symmetric tensor\n$$\nF_{i_1\\ldots i_n}=\\sum_{j_1,\\ldots,j_n, k_1\\ldots,k_n=1}^{n}\n\\epsilon_{j_1\\ldots j_n} \\epsilon_{k_1\\ldots k_n} T_{i_1,j_1,k_1}\\cdots T_{i_n,j_n,k_n}\\ .\n$$\nFor symmetric $T$'s, such as Will's diagonal tensor $D_{ijk}=\\delta_{ij}\\delta_{ik}$, this is the Hessian covariant.\nJason's idea can thus be rephrased as finding an $SL_n$-invariant $I=I(F)$ of $n$-ary $n$-ics $F$\nand proposing $I(\\Gamma(T))$ as an answer to Will's question. The main issue is to show $I(\\Gamma(D))\\neq 0$.\nNow $\\Gamma(D)$ is the form $F(x_1,\\ldots,x_n)=n!\\ x_1 x_2\\cdots x_n$.</p>\n\n<p>For $n$ even, the smallest degree for an invariant of $n$-ary $n$-ics $F(x_1,\\ldots,x_n)$ is $n$ (unique up to scale). Its symbolic form is\n$$\n\\mathcal{S_n}=(a^{(1)} a^{(2)}\\cdots a^{(n)})^n\n$$\nwhere $a^{(1)}=(a_{1}^{(1)},a_{2}^{(1)}\\ldots,a_{n}^{(1)}),\\ldots,a^{(n)}=(a_{1}^{(n)},a_{2}^{(n)}\\ldots,a_{n}^{(n)})$\nare $n$ series of $n$ variables. Let $I_n=\\Psi(\\mathcal{S_n})$ denote this invariant. One can also write\n$$\nI_n(F)=\\epsilon_{i_{1,1}\\ldots i_{1,n}}\\epsilon_{i_{2,1}\\ldots i_{2,n}}\\cdots\\epsilon_{i_{n,1}\\ldots i_{n,n}}\n\\ F_{i_{1,1}\\ldots i_{n,1}}F_{i_{1,2}\\ldots i_{n,2}}\\cdots\nF_{i_{1,n}\\ldots i_{n,n}}\n$$\nwith summation over indices understood.\nThe evaluation of $I_n(\\Gamma(D))$ is a multiple of the number of row-even minus the number of row-odd $n\\times n$\nLatin squares. Via work of Huang, Rota and Janssen showing that the result is nonzero is equivalent to the (widely open) Alon-Tarsi conjecture.</p>\n\n<p><strong>The following may be new (if not please let me know):</strong>\nFor $n$ odd, the smallest degree for an invariant of $F(x_1,\\ldots,x_n)$ is $n+1$ (unique up to scale). Its symbolic form is\n$$\n\\mathcal{S}=(a^{(2)} a^{(3)}\\cdots a^{(n+1)})(a^{(1)} a^{(3)}\\cdots a^{(n+1)})\\cdots(a^{(1)} a^{(2)}\\cdots a^{(n)})\\ .\n$$\nAgain let me define $I_n=\\Psi(\\mathcal{S_n})$. This is nonzero because the polynomial in the $a^{(i)}$ is not only multihomogeneous\nof multidegree $[n^{n+1}]$ but also multisymmetric. For instance the transposition $a^{(1)}\\leftrightarrow a^{(2)}$ exchanges the first\ntwo bracket factors while the other ones pick up a factor $(-1)$. This gives $(-1)^{n-1}=1$ since $n$ is odd. Other elementary\ntranspositions $a^{(i)}\\leftrightarrow a^{(i+1)}$ can be treated in the same way.\nOf course $I_3$ is Aronhold's degree 4 invariant. I don't know if $I_5$ could be of use for the study of quintic threefolds.\nI boldly (or foolishly?) make the following conjecture which I think is a natural analog of the Alon-Tarsi conjecture in the odd case.</p>\n\n<p><strong>Conjecture:</strong> For $n$ odd, $I_n$ does not vanish on $F(x_1,\\ldots,x_n)=x_1 x_2\\cdots x_n$.</p>\n\n<p>Here is a proof for $n=3$. The Aronhold invariant of degree 4 is the defining equation for the Veronese secant variety\n$\\sigma_{3}(v_3(\\mathbb{P}^2))$.\nHence, it does not vanish on forms with border rank $&gt;3$.\nThe border rank of the form $x_1 x_2\\cdots x_n$ is exactly $2^{n-1}$ as shown by Oeding in <a href=\"https://arxiv.org/abs/1608.02530\" rel=\"nofollow noreferrer\">this article</a>. The case $n=3$ relevant to the present proof was shown earlier by\n<a href=\"https://arxiv.org/abs/0901.0487\" rel=\"nofollow noreferrer\">Landsberg and Teitler</a>. QED.</p>\n\n<p>A final remark regarding the endgame in Jason's argument with a view to producing an explicit invariant of moderate degree: it follows from what I mentioned above that saying that any homogeneous multisymmetric\nfunction in $M_{[p^q]}^{n}\\simeq S^q(S^p(\\mathbb{C}^n))$ can be expressed as a polynomial in the elementary ones is equivalent\nto the Hermite-Hadamard-Howe map $H_{p,q}$ being surjective and this is known for $p$ large with an explicit bound due to Brion (see\nthe review by Landsberg I mentioned). The problem I was discussing is closer to the $p=q$ case related to the Foulkes-Howe conjecture.\nThe latter is known to be false, but not by much! (see <a href=\"https://arxiv.org/abs/1509.03944\" rel=\"nofollow noreferrer\">this article</a>).</p>\n\n<p>Edit: I learned from C. Ikenmeyer that the conjecture about the nonvanishing of $I_n$ for $n$ odd is not new and is stated in Remark 3.26 in\n<a href=\"https://arxiv.org/abs/1511.02927\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1511.02927</a>\nHowever, I noticed recently that this is equivalent to the Alon-Tarsi conjecture for $n$ even. Indeed, the above invariants $I_n$ satisfy the following identity:\n$$\nI_n(F(x_1,\\ldots,x_{n-1})x_n)=(-1)^{\\frac{n}{2}}\\frac{n!}{n^n}\\ I_{n-1}(F)\n$$\nfor all $n$ even and all forms $F$ which only depend on the first $n-1$ variables.</p>\n" }, { "answer_id": 259140, "score": 5, "text": "<p><B>Edit.</B> There was a mistake in the original formulation below of Theorem 2.4, p. 134 of Gelfand-Kapranov-Zelevinsky (now corrected). That theorem only applies after passing to the standard open affines. The upshot is that we know that there exists some integer $d_0&gt;0$ such that the determinant $D_{A,k}^{2d}$ gives a $\\textbf{SL}(A)\\times \\textbf{SL}(B)\\times \\textbf{SL}(C)$-invariant polynomial that is nonzero on Will Sawin's tensor for all $d\\geq d_0$. However, it is not immediately clear what exponent $d_0$ is necessary. An upper bound does follow from the Castelnuovo-Mumford regularity of the symmetric power $(\\mathbb{P}(A))^n//\\mathfrak{S}_n$ as a subvariety of $\\mathbb{P}\\text{Sym}^n_k(A)$, and there are bounds on this by Gotzmann. However, the Gotzmann bound is certainly too large for any effective computations.</p>\n\n<p><B>Original post.</B><br>\nI am just writing up my comments above as an answer. Let $k$ be a commutative, unital ring. Let $A$ be a free $k$-module of rank $n$. Denote by $\\mathbb{P}(A)$ the Proj of the graded $k$-algebra $$S^\\bullet_k(A^\\vee) = \\bigoplus_{d\\geq 0} \\text{Sym}^d_k(A^\\vee).$$<br>\nFor every integer $m\\geq 0$, denote by $(S^\\bullet_k(A^\\vee))^{\\otimes m}$ the $m$-fold tensor product $k$-algebra, $$(S^\\bullet_k(A^\\vee))^{\\otimes m} = S^\\bullet_k(A^\\vee)\\otimes_k \\cdots \\otimes_k S^\\bullet_k(A^\\vee).$$ This is naturally a $\\mathbb{Z}^m$-graded algebra, and it has a natural action of the symmetric group $\\mathfrak{S}_m$. Denote by $(S^\\bullet_k(A^\\vee))^{\\otimes m}_{\\langle 1,\\dots,1 \\rangle}$ the $\\mathbb{Z}$-graded, $\\mathfrak{S}_m$-fixed subalgebra, $$S^\\bullet_k(A^\\vee))^{\\otimes m}_{\\langle 1,\\dots, 1 \\rangle} = \\bigoplus_{d\\geq 0} \\text{Sym}^d_k(A^\\vee)\\otimes_k \\cdots \\otimes_k \\text{Sym}^d_k(A^\\vee).$$\nNow assume that $m!$ is invertible in $k$, so that $\\text{Sym}^m_k(A)^\\vee$ is canonically isomorphic to $\\text{Sym}^m_k(A^\\vee)$ as $k$-vector spaces. Then the transpose of the $m$-multilinear $k$-module homomorphism, $$A^m \\to \\text{Sym}^m_k(A), \\ \\ (a_1,\\dots,a_m)\\mapsto a_1\\cdots a_m,$$ induces a $k$-algebra homomorphism, $$S_k^\\bullet(\\text{Sym}^m_k(A^\\vee)) \\to (S^\\bullet_k(A^\\vee))^{\\otimes m}_{\\langle 1,\\dots,1 \\rangle}.$$ This, in turn, defines a $k$-morphism, $$\\gamma': (\\mathbb{P}(A))^m\\to \\mathbb{P}(\\text{Sym}^m_k(A)),\\ \\ ([a_1],\\dots,[a_m]) \\mapsto [a_1\\cdots a_m]$$ where $(\\mathbb{P}(A))^m$ is shorthand for the $m$-fold self-fiber product $\\mathbb{P}(A)\\times_{\\text{Spec}\\ k}\\cdots \\times_{\\text{Spec}\\ k}\\mathbb{P}(A)$. The morphism $\\gamma'$ is invariant for the natural action of the symmetric group $\\mathfrak{S}_m$ on $(\\mathbb{P}(A))^m$. Thus, $\\gamma'$ factors through the quotient of this group action. Denote by $\\gamma$ the induced $k$-morphism, $$\\gamma: (\\mathbb{P}(A))^m//\\mathfrak{S}_m \\to \\mathbb{P}(\\text{Sym}^m_k(A)).$$</p>\n\n<p>It is a classical fact that $\\gamma$ is a closed immersion. This is explicitly stated and proved in Proposition 2.1 and Theorem 2.2, Chapter 4, pp. 132-133 of Gelfand-Kapranov-Zelevinsky. In fact, Gelfand-Kapranov-Zelevinsky prove much more: for every nonzero $x_i\\in A^\\vee$, denoting by $\\gamma_m(x_i)\\in \\text{Sym}^m_k(A^\\vee)$ the associated $m^{\\text{th}}$ power, then the induced $k$-algebra homomorphism of affine opens, $$S_k^\\bullet(\\text{Sym}^m_k(A^\\vee))[1/\\gamma_m(x_i)] \\to [(S_k^\\bullet(A^\\vee))^{\\otimes m}_{\\langle 1,\\dots,1 \\rangle}]^{\\mathfrak{S}_m}[1/(x_{1,i}\\cdots x_{m,i})],$$ is a surjective $k$-algebra homomorphism. This is essentially Theorem 2.4 on p. 134. From this it follows that for every $\\mathfrak{S}_m$-invariant, homogeneous element $D$ of degree $r&gt;0$ in $(S_k^\\bullet(A^\\vee))^{\\otimes m}_{(r,\\dots,r)}$, there exists an integer $e_0$ such that for every $e\\geq e_0$, $D$ times the image of $\\text{Sym}^e_k(\\text{Sym}^m_k(A^\\vee))$ is in the image of $\\text{Sym}^{e+r}_k(\\text{Sym}^m_k(A^\\vee))$. Thus there exists an integer $d&gt;0$ such that $D^d$ is in the image of $\\text{Sym}^{dr}_k(\\text{Sym}^m_k(A^\\vee))$.</p>\n\n<p>In particular, for $m=n$ (assuming that $n!$ is invertible in $k$), the determinant map, $$\\text{det}_{A,k}: A^n \\to \\bigwedge^n_k(A),$$ defines an element $D_{A,k} \\in (S_k^\\bullet(A^\\vee))^{\\otimes n}_{(1,\\dots,1)}$. This element is <B>not</B> $\\mathfrak{S}_m$-invariant. However, the square, $D_{A,k}^2$ is $\\mathfrak{S}_m$-invariant. Thus, $D_{A,k}^{2d}$ is in the image $I_{2d}$ of $\\text{Sym}_k^{2d}(\\text{Sym}_k^n(A^\\vee))$ for some integer $d&gt;0$.</p>\n\n<p>Finally, if we assume that $k$ is a field of characteristic $0$, then $\\textbf{SL}(A)$ is linearly reductive. Thus, the induced $k$-linear map, $$\\text{Sym}^{2d}_k(\\text{Sym}^m_k(A^\\vee))^{\\textbf{SL}(A)} \\to I_{2d}^{\\textbf{SL}(A)},$$\nis surjective. In particular, the $\\textbf{SL}(A)$-invariant element $D_{A,k}^{2d}$ is in the image of $\\left( \\text{Sym}_k^{2d}(\\text{Sym}_k^n(A^\\vee)) \\right)^{\\textbf{SL}(A)}$. So there exists some $\\textbf{SL}(A)$-invariant element $\\Delta_{A,k}$ mappsing to $D_{A,k}^{2d}$.</p>\n\n<p>Now put this together with the tensor operation from the comments. For $k$-vector spaces $A$, $B$, and $C$ of dimension $n$, for every $k$-linear map, $$t:A^\\vee \\to \\text{Hom}_k(B,C),$$ there is an associated element $$\\widetilde{t} \\in \\text{Sym}^n(A) \\otimes_k \\text{Hom}_k(\\bigwedge^n_k(B),\\bigwedge^n_k(C)),$$ giving the polynomial map, $$ A^\\vee\\to \\text{Hom}_k(\\bigwedge^n_k(B),\\bigwedge^n_k(C)), \\ \\ a \\mapsto \\text{det}(t(a)).$$ The associated map $$\\text{Hom}_k(A^\\vee,\\text{Hom}_k(B,C))\\to \\text{Sym}^n(A)\\otimes_k \\text{Hom}_k(\\bigwedge^n_k(B),\\bigwedge^n_k(C)), \\ \\ t\\mapsto \\widetilde{t},$$ is a degree $n$, homogeneous $k$-polynomial map that is equivariant for the natural actions of $\\textbf{SL}(A)\\times \\textbf{SL}(B)\\times \\textbf{SL}(C)$. Via the duality of $\\text{Sym}^n(A)$ and $\\text{Sym}^n(A^\\vee)$, the polynomial $\\Delta_{A,k}$ determines a $\\textbf{SL}(A)\\times \\textbf{SL}(B)\\times \\textbf{SL}(C)$-invariant homogeneous polynomial of degree $2dn$ on $\\text{Hom}_k(A^\\vee,\\text{Hom}_k(B,C))$. </p>\n\n<p>For the specific tensor $t$ that you wrote down, with respect to appropriate bases for $A$, $B$, and $C$, for every $(a_1,\\dots,a_n)\\in A$, $t(a_1,\\dots,a_n)$ is the linear transformation whose matrix representative is the diagonal $n\\times n$ matrix with entries $(a_1,\\dots,a_n)$. Thus, $\\widetilde{t}$ equals $(a_1\\cdots a_n)\\otimes (b_1\\wedge \\dots \\wedge b_n)\\otimes (c_1\\wedge \\dots c_n)$. In particular, since $a_1\\cdots a_n$ is in the image of $\\gamma$, $\\Delta_{A,k}$ restricts on $\\widetilde{t}$ as a power of the determinant. Since $a_1\\cdots a_n$ is a product of $n$ linearly independent linear polynomials, the determinant is nonzero. Thus, $\\Delta_{A,k}$ is a $\\textbf{SL}(A)\\times \\textbf{SL}(B)\\times \\textbf{SL}(C)$-invariant, homogeneous polynomial of degree $2n$ that is nonzero on your tensor. </p>\n" }, { "answer_id": 259195, "score": 9, "text": "<p>Let me start with some remarks about the classical symbolic method (without which one cannot understand 19th century invariant theory)\nand multisymmetric functions.\nI will use an example first. Take four series of three variables $a=(a_1,a_2,a_3)$, $b=(b_1,b_2,b_3)$,\n$c=(c_1,c_2,c_3)$ and $d=(d_1,d_2,d_3)$.\nNow define the polynomial $\\mathcal{A}$ in these 12 indeterminates given by\n$$\n\\mathcal{A}=(bcd)(acd)(abd)(abc)\n$$\nwhere we used a shorthand notation for determinants\n$$\n(bcd)=\\left|\n\\begin{array}{ccc}\nb_1 &amp; b_2 &amp; b_3 \\\\\nc_1 &amp; c_2 &amp; c_3 \\\\\nd_1 &amp; d_2 &amp; d_3 \\\\\n\\end{array}\n\\right|\n$$\nand similarly for the other \"bracket factors\" $(acd)$, etc.\nThis is an example of multisymmetric function in the \"letters\" $a,b,c,d$ which is multihomogeneous of multidegree $[3,3,3,3]$.\nLet me write $M_{[3,3,3,3]}^3$ for the space of such multisymmetric functions. The subscript is the multidegree while the superscript refers\nto the number of variables in each series. \nThe symmetric group acting is $\\mathfrak{S}_4$ which permutes these series. If each series had one variable in it instead of\nthree we would be in the familiar realm of the theory of ordinary symmetric functions. The analogous theory of multisymmetric\nfunctions is much more complicated and has been studied classically by Euler, Lagrange, Poisson, Schl&auml;fli, Brill, Gordan, Junker\nand more recently by Angeniol, Dalbec, Briand, Vaccarino, Domokos and others.\nA nice yet old reference on this is the book on elimination theory by Faa di Bruno I mentioned in <a href=\"https://mathoverflow.net/questions/117415/old-books-still-used/117417#117417\">this MO answer</a>.</p>\n\n<p>Consider now a generic ternary cubic form $F=F(x_1,x_2,x_3)$. One can write it as\n$$\nF(x)=\\sum_{i_1,i_2,i_3=1}^3 F_{i_1,i_2,i_3} x_{i_1} x_{i_2} x_{i_3}\n$$\nwhere $(F_{i_1,i_2,i_3})$ is a symmetric tensor (by which I mean a \"matrix\" rather than a Bourbaki tensor).\nThis form lives in the symmetric power $S^3(\\mathbb{C}^3)$ but I will be (intentionally) sloppy and write $S^3$ instead.\nThe space of degree 4 homogeneous polynomials in the coefficients of $F$ will be denoted by $S^4(S^3)$.</p>\n\n<p>Let me define a linear map $\\Phi:S^4(S^3)\\rightarrow M_{[3,3,3,3]}^3$ as follows.\nGiven a degree four homogeneous polynomial $C(F)$ first construct its full polarization\n$$\n\\widetilde{C}(F_1,F_2,F_3,F_4)=\\frac{1}{4!} \\frac{\\partial^4}{\\partial t_1\\partial t_2\\partial t_3\\partial t_4}\nC(t_1F_1+t_2 F_2+t_3 F_3+t_4 F_4)\n$$ \nwhich satisfies $\\widetilde{C}(F,F,F,F)=C(F)$. \nNow set $F_1(x)=a_{x}^3$ where $a_x$ is the classical notation for the linear form $a_1x_1+a_2x_2+a_3x_3$ and similarly\nspecialize to the cubes of linear forms $F_2(x)=b_{x}^3$, $F_3(x)=c_{x}^3$ and $F_4(x)=d_{x}^3$.\nFinally, define $\\Phi(C)=\\widetilde{C}(F_1,F_2,F_3,F_4)$.</p>\n\n<p>I will also define another linear map $\\Psi: M_{[3,3,3,3]}^3\\rightarrow S^4(S^3)$.\nLet $\\mathcal{S}=\\mathcal{S}(a,b,c,d)$ be a multisymmetric function in $M_{[3,3,3,3]}^3$.\nWe define the differential operator\n$$\n\\mathcal{D}=\\frac{1}{3!^4} \nF\\left(\\frac{\\partial}{\\partial a_1},\\frac{\\partial}{\\partial a_2},\\frac{\\partial}{\\partial a_3}\\right)\nF\\left(\\frac{\\partial}{\\partial b_1},\\frac{\\partial}{\\partial b_2},\\frac{\\partial}{\\partial b_3}\\right)\n$$\n$$\nF\\left(\\frac{\\partial}{\\partial c_1},\\frac{\\partial}{\\partial c_2},\\frac{\\partial}{\\partial c_3}\\right)\nF\\left(\\frac{\\partial}{\\partial d_1},\\frac{\\partial}{\\partial d_2},\\frac{\\partial}{\\partial d_3}\\right)\n$$\nand let\n$\\Psi(\\mathcal{S})=\\mathcal{D}\\ \\mathcal{S}$, i.e., the result of applying the differential operator $\\mathcal{D}$\nto the polynomial\n$\\mathcal{S}$.</p>\n\n<p>It is not hard to see that $\\Phi$ and $\\Psi$ are inverses of one another.\nIn the particular case of $\\mathcal{A}$ above, the image $\\Psi(\\mathcal{A})$ is Aronhold's degree four $SL_3$ invariant of plane cubics\nand $\\mathcal{A}$ itself is called the (German) symbolic form of that invariant. The classics usually omitted \"$\\Psi(\\cdots)$\" but they \nknew perfectly well what they were doing. Also note that this formalism, in an essentially equivalent form, was invented earlier by Cayley\nas I briefly explained in Section 1.11.4 of my article\n<a href=\"https://arxiv.org/abs/1604.05259\" rel=\"nofollow noreferrer\">\"A Second-Quantized Kolmogorov-Chentsov Theorem\"</a>.\nOne can easily write $\\Psi(\\mathcal{A})$ applied to $F$, in \"physicsy\" tensor contraction notation as\n$$\n\\epsilon_{i_{1} i_{2} i_{3}}\\epsilon_{i_{4} i_{5} i_{6}}\\epsilon_{i_{7} i_{8} i_{9}}\\epsilon_{i_{10} i_{11} i_{12}}\nF_{i_{4} i_{7} i_{10}} F_{i_{1} i_{8} i_{11}} F_{i_{2} i_{5} i_{12}} F_{i_{3} i_{6} i_{9}}\n$$\nwith summation over all twelve indices in the set $\\{1,2,3\\}$. I used the standard physics notation for the Levi-Civita tensor\n$\\epsilon_{i_1 i_2 i_3}$ equal to zero if some indices are equal and otherwise to the sign of the permutation $i_1,i_2,i_3$\nof $1,2,3$. This notation is consistent with the one I used in my previous answers to these MO questions:\n<a href=\"https://mathoverflow.net/questions/121715/invariant-polynomials-for-a-product-of-algebraic-groups\">Q1</a>,\n<a href=\"https://mathoverflow.net/questions/257709/ring-of-invariants-of-operatornamesl-6-acting-on-lambda3-mathbb-c6\">Q2</a>,\nand <a href=\"https://mathoverflow.net/questions/255492/how-to-constructively-combinatorially-prove-schur-weyl-duality\">Q3</a>.\nAt bottom, this is an intrinsically diagrammatic language and if the reader had trouble following what I said so far, it's probably because\nI did not draw any pictures or \"finite-dimensional Feynman diagrams\" (I don't know how to do this on MO). To see these pictures,\nplease have a look at <a href=\"https://arxiv.org/abs/1402.2669\" rel=\"nofollow noreferrer\">this article</a> about an invariant of cubic threefolds or Section 2\nof <a href=\"https://arxiv.org/abs/0904.1734\" rel=\"nofollow noreferrer\">this one</a> mostly on binary forms.\nIn a nutshell, the above contraction can be encoded in a bipartite graph with 4 $\\epsilon$-vertices and 4 $F$-vertices, all of degree three.</p>\n\n<p>Of course the above formalism generalizes with no difficulty (except managing notations) to plethysms of symmetric powers\n$S^p(S^q(\\mathbb{C}^n))$ and similar $\\Phi$, $\\Psi$ maps giving the isomorphism with the relevant space of multisymmetric functions, i.e.,\n$M_{[q,q,\\ldots,q]}^{n}$ (with $q$ repeated $p$ times) or in simpler notation $M_{[q^p]}^{n}$.\nThere is a canonical $GL_n$-equivariant map\n$$\nH_{p,q}:S^p(S^q(\\mathbb{C}^n))\\rightarrow S^q(S^p(\\mathbb{C}^n))\n$$ \nwith \"H\" standing for Hermite, Hadamard or Howe. I will not define it here but it's the first thing that would come to mind\nif asked to produce such a map\n(see <a href=\"https://arxiv.org/abs/1305.7387\" rel=\"nofollow noreferrer\">this review</a> by Landsberg for a detailed discussion in the wider context of geometric complexity theory).\nIt is not hard to see that a homogeneous multisymmetric function in $M_{[p^q]}^{n}\\simeq S^q(S^p(\\mathbb{C}^n))$ is in the image\nof $H_{p,q}$ iff it is a polynomial in the homogeneous elementary multisymmetric functions which belong to $M_{[1^q]}^{n}$.\nIn the context of the Aronhold invariant above, these elementary functions are the coefficients of the polynomial\n$$\nP(x_1,x_2,x_3)=a_x b_x c_x d_x\n$$ \nand similarly in general.</p>\n\n<p>Finally, I can start discussing Will's excellent question. Let $T_{i_1,i_2,i_3}$ denote the tensor of interest with the action of\n$(g,h,k)\\in (SL_n)^3$ given by\n$$\n[(g,h,k)\\cdot T]_{i_1,i_2,i_3}=\\sum_{j_1,j_2,j_3=1}^{n} g_{i_1 j_1} h_{i_2 j_2} k_{i_3 j_3} T_{j_1,j_2,j_3}\\ .\n$$\nIf Will's tensor had an even number $r$ of indices with the action of $(SL_n)^r$ the answer would be trivial.\nTake $n$ copies of $T$, contract all first indices of these copies with one $\\epsilon$,\nthen all second indices with another $\\epsilon$, etc. For $r=3$ this tentative invariant vanishes identically.\nOne can however keep the first indices free, and this gives a map $T\\mapsto \\Gamma(T)=F$ producing an $n$-ary $n$-ic\n$F(x_1,\\ldots,x_n)$ corresponding to the symmetric tensor\n$$\nF_{i_1\\ldots i_n}=\\sum_{j_1,\\ldots,j_n, k_1\\ldots,k_n=1}^{n}\n\\epsilon_{j_1\\ldots j_n} \\epsilon_{k_1\\ldots k_n} T_{i_1,j_1,k_1}\\cdots T_{i_n,j_n,k_n}\\ .\n$$\nFor symmetric $T$'s, such as Will's diagonal tensor $D_{ijk}=\\delta_{ij}\\delta_{ik}$, this is the Hessian covariant.\nJason's idea can thus be rephrased as finding an $SL_n$-invariant $I=I(F)$ of $n$-ary $n$-ics $F$\nand proposing $I(\\Gamma(T))$ as an answer to Will's question. The main issue is to show $I(\\Gamma(D))\\neq 0$.\nNow $\\Gamma(D)$ is the form $F(x_1,\\ldots,x_n)=n!\\ x_1 x_2\\cdots x_n$.</p>\n\n<p>For $n$ even, the smallest degree for an invariant of $n$-ary $n$-ics $F(x_1,\\ldots,x_n)$ is $n$ (unique up to scale). Its symbolic form is\n$$\n\\mathcal{S_n}=(a^{(1)} a^{(2)}\\cdots a^{(n)})^n\n$$\nwhere $a^{(1)}=(a_{1}^{(1)},a_{2}^{(1)}\\ldots,a_{n}^{(1)}),\\ldots,a^{(n)}=(a_{1}^{(n)},a_{2}^{(n)}\\ldots,a_{n}^{(n)})$\nare $n$ series of $n$ variables. Let $I_n=\\Psi(\\mathcal{S_n})$ denote this invariant. One can also write\n$$\nI_n(F)=\\epsilon_{i_{1,1}\\ldots i_{1,n}}\\epsilon_{i_{2,1}\\ldots i_{2,n}}\\cdots\\epsilon_{i_{n,1}\\ldots i_{n,n}}\n\\ F_{i_{1,1}\\ldots i_{n,1}}F_{i_{1,2}\\ldots i_{n,2}}\\cdots\nF_{i_{1,n}\\ldots i_{n,n}}\n$$\nwith summation over indices understood.\nThe evaluation of $I_n(\\Gamma(D))$ is a multiple of the number of row-even minus the number of row-odd $n\\times n$\nLatin squares. Via work of Huang, Rota and Janssen showing that the result is nonzero is equivalent to the (widely open) Alon-Tarsi conjecture.</p>\n\n<p><strong>The following may be new (if not please let me know):</strong>\nFor $n$ odd, the smallest degree for an invariant of $F(x_1,\\ldots,x_n)$ is $n+1$ (unique up to scale). Its symbolic form is\n$$\n\\mathcal{S}=(a^{(2)} a^{(3)}\\cdots a^{(n+1)})(a^{(1)} a^{(3)}\\cdots a^{(n+1)})\\cdots(a^{(1)} a^{(2)}\\cdots a^{(n)})\\ .\n$$\nAgain let me define $I_n=\\Psi(\\mathcal{S_n})$. This is nonzero because the polynomial in the $a^{(i)}$ is not only multihomogeneous\nof multidegree $[n^{n+1}]$ but also multisymmetric. For instance the transposition $a^{(1)}\\leftrightarrow a^{(2)}$ exchanges the first\ntwo bracket factors while the other ones pick up a factor $(-1)$. This gives $(-1)^{n-1}=1$ since $n$ is odd. Other elementary\ntranspositions $a^{(i)}\\leftrightarrow a^{(i+1)}$ can be treated in the same way.\nOf course $I_3$ is Aronhold's degree 4 invariant. I don't know if $I_5$ could be of use for the study of quintic threefolds.\nI boldly (or foolishly?) make the following conjecture which I think is a natural analog of the Alon-Tarsi conjecture in the odd case.</p>\n\n<p><strong>Conjecture:</strong> For $n$ odd, $I_n$ does not vanish on $F(x_1,\\ldots,x_n)=x_1 x_2\\cdots x_n$.</p>\n\n<p>Here is a proof for $n=3$. The Aronhold invariant of degree 4 is the defining equation for the Veronese secant variety\n$\\sigma_{3}(v_3(\\mathbb{P}^2))$.\nHence, it does not vanish on forms with border rank $&gt;3$.\nThe border rank of the form $x_1 x_2\\cdots x_n$ is exactly $2^{n-1}$ as shown by Oeding in <a href=\"https://arxiv.org/abs/1608.02530\" rel=\"nofollow noreferrer\">this article</a>. The case $n=3$ relevant to the present proof was shown earlier by\n<a href=\"https://arxiv.org/abs/0901.0487\" rel=\"nofollow noreferrer\">Landsberg and Teitler</a>. QED.</p>\n\n<p>A final remark regarding the endgame in Jason's argument with a view to producing an explicit invariant of moderate degree: it follows from what I mentioned above that saying that any homogeneous multisymmetric\nfunction in $M_{[p^q]}^{n}\\simeq S^q(S^p(\\mathbb{C}^n))$ can be expressed as a polynomial in the elementary ones is equivalent\nto the Hermite-Hadamard-Howe map $H_{p,q}$ being surjective and this is known for $p$ large with an explicit bound due to Brion (see\nthe review by Landsberg I mentioned). The problem I was discussing is closer to the $p=q$ case related to the Foulkes-Howe conjecture.\nThe latter is known to be false, but not by much! (see <a href=\"https://arxiv.org/abs/1509.03944\" rel=\"nofollow noreferrer\">this article</a>).</p>\n\n<p>Edit: I learned from C. Ikenmeyer that the conjecture about the nonvanishing of $I_n$ for $n$ odd is not new and is stated in Remark 3.26 in\n<a href=\"https://arxiv.org/abs/1511.02927\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1511.02927</a>\nHowever, I noticed recently that this is equivalent to the Alon-Tarsi conjecture for $n$ even. Indeed, the above invariants $I_n$ satisfy the following identity:\n$$\nI_n(F(x_1,\\ldots,x_{n-1})x_n)=(-1)^{\\frac{n}{2}}\\frac{n!}{n^n}\\ I_{n-1}(F)\n$$\nfor all $n$ even and all forms $F$ which only depend on the first $n-1$ variables.</p>\n" }, { "answer_id": 259140, "score": 5, "text": "<p><B>Edit.</B> There was a mistake in the original formulation below of Theorem 2.4, p. 134 of Gelfand-Kapranov-Zelevinsky (now corrected). That theorem only applies after passing to the standard open affines. The upshot is that we know that there exists some integer $d_0&gt;0$ such that the determinant $D_{A,k}^{2d}$ gives a $\\textbf{SL}(A)\\times \\textbf{SL}(B)\\times \\textbf{SL}(C)$-invariant polynomial that is nonzero on Will Sawin's tensor for all $d\\geq d_0$. However, it is not immediately clear what exponent $d_0$ is necessary. An upper bound does follow from the Castelnuovo-Mumford regularity of the symmetric power $(\\mathbb{P}(A))^n//\\mathfrak{S}_n$ as a subvariety of $\\mathbb{P}\\text{Sym}^n_k(A)$, and there are bounds on this by Gotzmann. However, the Gotzmann bound is certainly too large for any effective computations.</p>\n\n<p><B>Original post.</B><br>\nI am just writing up my comments above as an answer. Let $k$ be a commutative, unital ring. Let $A$ be a free $k$-module of rank $n$. Denote by $\\mathbb{P}(A)$ the Proj of the graded $k$-algebra $$S^\\bullet_k(A^\\vee) = \\bigoplus_{d\\geq 0} \\text{Sym}^d_k(A^\\vee).$$<br>\nFor every integer $m\\geq 0$, denote by $(S^\\bullet_k(A^\\vee))^{\\otimes m}$ the $m$-fold tensor product $k$-algebra, $$(S^\\bullet_k(A^\\vee))^{\\otimes m} = S^\\bullet_k(A^\\vee)\\otimes_k \\cdots \\otimes_k S^\\bullet_k(A^\\vee).$$ This is naturally a $\\mathbb{Z}^m$-graded algebra, and it has a natural action of the symmetric group $\\mathfrak{S}_m$. Denote by $(S^\\bullet_k(A^\\vee))^{\\otimes m}_{\\langle 1,\\dots,1 \\rangle}$ the $\\mathbb{Z}$-graded, $\\mathfrak{S}_m$-fixed subalgebra, $$S^\\bullet_k(A^\\vee))^{\\otimes m}_{\\langle 1,\\dots, 1 \\rangle} = \\bigoplus_{d\\geq 0} \\text{Sym}^d_k(A^\\vee)\\otimes_k \\cdots \\otimes_k \\text{Sym}^d_k(A^\\vee).$$\nNow assume that $m!$ is invertible in $k$, so that $\\text{Sym}^m_k(A)^\\vee$ is canonically isomorphic to $\\text{Sym}^m_k(A^\\vee)$ as $k$-vector spaces. Then the transpose of the $m$-multilinear $k$-module homomorphism, $$A^m \\to \\text{Sym}^m_k(A), \\ \\ (a_1,\\dots,a_m)\\mapsto a_1\\cdots a_m,$$ induces a $k$-algebra homomorphism, $$S_k^\\bullet(\\text{Sym}^m_k(A^\\vee)) \\to (S^\\bullet_k(A^\\vee))^{\\otimes m}_{\\langle 1,\\dots,1 \\rangle}.$$ This, in turn, defines a $k$-morphism, $$\\gamma': (\\mathbb{P}(A))^m\\to \\mathbb{P}(\\text{Sym}^m_k(A)),\\ \\ ([a_1],\\dots,[a_m]) \\mapsto [a_1\\cdots a_m]$$ where $(\\mathbb{P}(A))^m$ is shorthand for the $m$-fold self-fiber product $\\mathbb{P}(A)\\times_{\\text{Spec}\\ k}\\cdots \\times_{\\text{Spec}\\ k}\\mathbb{P}(A)$. The morphism $\\gamma'$ is invariant for the natural action of the symmetric group $\\mathfrak{S}_m$ on $(\\mathbb{P}(A))^m$. Thus, $\\gamma'$ factors through the quotient of this group action. Denote by $\\gamma$ the induced $k$-morphism, $$\\gamma: (\\mathbb{P}(A))^m//\\mathfrak{S}_m \\to \\mathbb{P}(\\text{Sym}^m_k(A)).$$</p>\n\n<p>It is a classical fact that $\\gamma$ is a closed immersion. This is explicitly stated and proved in Proposition 2.1 and Theorem 2.2, Chapter 4, pp. 132-133 of Gelfand-Kapranov-Zelevinsky. In fact, Gelfand-Kapranov-Zelevinsky prove much more: for every nonzero $x_i\\in A^\\vee$, denoting by $\\gamma_m(x_i)\\in \\text{Sym}^m_k(A^\\vee)$ the associated $m^{\\text{th}}$ power, then the induced $k$-algebra homomorphism of affine opens, $$S_k^\\bullet(\\text{Sym}^m_k(A^\\vee))[1/\\gamma_m(x_i)] \\to [(S_k^\\bullet(A^\\vee))^{\\otimes m}_{\\langle 1,\\dots,1 \\rangle}]^{\\mathfrak{S}_m}[1/(x_{1,i}\\cdots x_{m,i})],$$ is a surjective $k$-algebra homomorphism. This is essentially Theorem 2.4 on p. 134. From this it follows that for every $\\mathfrak{S}_m$-invariant, homogeneous element $D$ of degree $r&gt;0$ in $(S_k^\\bullet(A^\\vee))^{\\otimes m}_{(r,\\dots,r)}$, there exists an integer $e_0$ such that for every $e\\geq e_0$, $D$ times the image of $\\text{Sym}^e_k(\\text{Sym}^m_k(A^\\vee))$ is in the image of $\\text{Sym}^{e+r}_k(\\text{Sym}^m_k(A^\\vee))$. Thus there exists an integer $d&gt;0$ such that $D^d$ is in the image of $\\text{Sym}^{dr}_k(\\text{Sym}^m_k(A^\\vee))$.</p>\n\n<p>In particular, for $m=n$ (assuming that $n!$ is invertible in $k$), the determinant map, $$\\text{det}_{A,k}: A^n \\to \\bigwedge^n_k(A),$$ defines an element $D_{A,k} \\in (S_k^\\bullet(A^\\vee))^{\\otimes n}_{(1,\\dots,1)}$. This element is <B>not</B> $\\mathfrak{S}_m$-invariant. However, the square, $D_{A,k}^2$ is $\\mathfrak{S}_m$-invariant. Thus, $D_{A,k}^{2d}$ is in the image $I_{2d}$ of $\\text{Sym}_k^{2d}(\\text{Sym}_k^n(A^\\vee))$ for some integer $d&gt;0$.</p>\n\n<p>Finally, if we assume that $k$ is a field of characteristic $0$, then $\\textbf{SL}(A)$ is linearly reductive. Thus, the induced $k$-linear map, $$\\text{Sym}^{2d}_k(\\text{Sym}^m_k(A^\\vee))^{\\textbf{SL}(A)} \\to I_{2d}^{\\textbf{SL}(A)},$$\nis surjective. In particular, the $\\textbf{SL}(A)$-invariant element $D_{A,k}^{2d}$ is in the image of $\\left( \\text{Sym}_k^{2d}(\\text{Sym}_k^n(A^\\vee)) \\right)^{\\textbf{SL}(A)}$. So there exists some $\\textbf{SL}(A)$-invariant element $\\Delta_{A,k}$ mappsing to $D_{A,k}^{2d}$.</p>\n\n<p>Now put this together with the tensor operation from the comments. For $k$-vector spaces $A$, $B$, and $C$ of dimension $n$, for every $k$-linear map, $$t:A^\\vee \\to \\text{Hom}_k(B,C),$$ there is an associated element $$\\widetilde{t} \\in \\text{Sym}^n(A) \\otimes_k \\text{Hom}_k(\\bigwedge^n_k(B),\\bigwedge^n_k(C)),$$ giving the polynomial map, $$ A^\\vee\\to \\text{Hom}_k(\\bigwedge^n_k(B),\\bigwedge^n_k(C)), \\ \\ a \\mapsto \\text{det}(t(a)).$$ The associated map $$\\text{Hom}_k(A^\\vee,\\text{Hom}_k(B,C))\\to \\text{Sym}^n(A)\\otimes_k \\text{Hom}_k(\\bigwedge^n_k(B),\\bigwedge^n_k(C)), \\ \\ t\\mapsto \\widetilde{t},$$ is a degree $n$, homogeneous $k$-polynomial map that is equivariant for the natural actions of $\\textbf{SL}(A)\\times \\textbf{SL}(B)\\times \\textbf{SL}(C)$. Via the duality of $\\text{Sym}^n(A)$ and $\\text{Sym}^n(A^\\vee)$, the polynomial $\\Delta_{A,k}$ determines a $\\textbf{SL}(A)\\times \\textbf{SL}(B)\\times \\textbf{SL}(C)$-invariant homogeneous polynomial of degree $2dn$ on $\\text{Hom}_k(A^\\vee,\\text{Hom}_k(B,C))$. </p>\n\n<p>For the specific tensor $t$ that you wrote down, with respect to appropriate bases for $A$, $B$, and $C$, for every $(a_1,\\dots,a_n)\\in A$, $t(a_1,\\dots,a_n)$ is the linear transformation whose matrix representative is the diagonal $n\\times n$ matrix with entries $(a_1,\\dots,a_n)$. Thus, $\\widetilde{t}$ equals $(a_1\\cdots a_n)\\otimes (b_1\\wedge \\dots \\wedge b_n)\\otimes (c_1\\wedge \\dots c_n)$. In particular, since $a_1\\cdots a_n$ is in the image of $\\gamma$, $\\Delta_{A,k}$ restricts on $\\widetilde{t}$ as a power of the determinant. Since $a_1\\cdots a_n$ is a product of $n$ linearly independent linear polynomials, the determinant is nonzero. Thus, $\\Delta_{A,k}$ is a $\\textbf{SL}(A)\\times \\textbf{SL}(B)\\times \\textbf{SL}(C)$-invariant, homogeneous polynomial of degree $2n$ that is nonzero on your tensor. </p>\n" }, { "answer_id": 259195, "score": 9, "text": "<p>Let me start with some remarks about the classical symbolic method (without which one cannot understand 19th century invariant theory)\nand multisymmetric functions.\nI will use an example first. Take four series of three variables $a=(a_1,a_2,a_3)$, $b=(b_1,b_2,b_3)$,\n$c=(c_1,c_2,c_3)$ and $d=(d_1,d_2,d_3)$.\nNow define the polynomial $\\mathcal{A}$ in these 12 indeterminates given by\n$$\n\\mathcal{A}=(bcd)(acd)(abd)(abc)\n$$\nwhere we used a shorthand notation for determinants\n$$\n(bcd)=\\left|\n\\begin{array}{ccc}\nb_1 &amp; b_2 &amp; b_3 \\\\\nc_1 &amp; c_2 &amp; c_3 \\\\\nd_1 &amp; d_2 &amp; d_3 \\\\\n\\end{array}\n\\right|\n$$\nand similarly for the other \"bracket factors\" $(acd)$, etc.\nThis is an example of multisymmetric function in the \"letters\" $a,b,c,d$ which is multihomogeneous of multidegree $[3,3,3,3]$.\nLet me write $M_{[3,3,3,3]}^3$ for the space of such multisymmetric functions. The subscript is the multidegree while the superscript refers\nto the number of variables in each series. \nThe symmetric group acting is $\\mathfrak{S}_4$ which permutes these series. If each series had one variable in it instead of\nthree we would be in the familiar realm of the theory of ordinary symmetric functions. The analogous theory of multisymmetric\nfunctions is much more complicated and has been studied classically by Euler, Lagrange, Poisson, Schl&auml;fli, Brill, Gordan, Junker\nand more recently by Angeniol, Dalbec, Briand, Vaccarino, Domokos and others.\nA nice yet old reference on this is the book on elimination theory by Faa di Bruno I mentioned in <a href=\"https://mathoverflow.net/questions/117415/old-books-still-used/117417#117417\">this MO answer</a>.</p>\n\n<p>Consider now a generic ternary cubic form $F=F(x_1,x_2,x_3)$. One can write it as\n$$\nF(x)=\\sum_{i_1,i_2,i_3=1}^3 F_{i_1,i_2,i_3} x_{i_1} x_{i_2} x_{i_3}\n$$\nwhere $(F_{i_1,i_2,i_3})$ is a symmetric tensor (by which I mean a \"matrix\" rather than a Bourbaki tensor).\nThis form lives in the symmetric power $S^3(\\mathbb{C}^3)$ but I will be (intentionally) sloppy and write $S^3$ instead.\nThe space of degree 4 homogeneous polynomials in the coefficients of $F$ will be denoted by $S^4(S^3)$.</p>\n\n<p>Let me define a linear map $\\Phi:S^4(S^3)\\rightarrow M_{[3,3,3,3]}^3$ as follows.\nGiven a degree four homogeneous polynomial $C(F)$ first construct its full polarization\n$$\n\\widetilde{C}(F_1,F_2,F_3,F_4)=\\frac{1}{4!} \\frac{\\partial^4}{\\partial t_1\\partial t_2\\partial t_3\\partial t_4}\nC(t_1F_1+t_2 F_2+t_3 F_3+t_4 F_4)\n$$ \nwhich satisfies $\\widetilde{C}(F,F,F,F)=C(F)$. \nNow set $F_1(x)=a_{x}^3$ where $a_x$ is the classical notation for the linear form $a_1x_1+a_2x_2+a_3x_3$ and similarly\nspecialize to the cubes of linear forms $F_2(x)=b_{x}^3$, $F_3(x)=c_{x}^3$ and $F_4(x)=d_{x}^3$.\nFinally, define $\\Phi(C)=\\widetilde{C}(F_1,F_2,F_3,F_4)$.</p>\n\n<p>I will also define another linear map $\\Psi: M_{[3,3,3,3]}^3\\rightarrow S^4(S^3)$.\nLet $\\mathcal{S}=\\mathcal{S}(a,b,c,d)$ be a multisymmetric function in $M_{[3,3,3,3]}^3$.\nWe define the differential operator\n$$\n\\mathcal{D}=\\frac{1}{3!^4} \nF\\left(\\frac{\\partial}{\\partial a_1},\\frac{\\partial}{\\partial a_2},\\frac{\\partial}{\\partial a_3}\\right)\nF\\left(\\frac{\\partial}{\\partial b_1},\\frac{\\partial}{\\partial b_2},\\frac{\\partial}{\\partial b_3}\\right)\n$$\n$$\nF\\left(\\frac{\\partial}{\\partial c_1},\\frac{\\partial}{\\partial c_2},\\frac{\\partial}{\\partial c_3}\\right)\nF\\left(\\frac{\\partial}{\\partial d_1},\\frac{\\partial}{\\partial d_2},\\frac{\\partial}{\\partial d_3}\\right)\n$$\nand let\n$\\Psi(\\mathcal{S})=\\mathcal{D}\\ \\mathcal{S}$, i.e., the result of applying the differential operator $\\mathcal{D}$\nto the polynomial\n$\\mathcal{S}$.</p>\n\n<p>It is not hard to see that $\\Phi$ and $\\Psi$ are inverses of one another.\nIn the particular case of $\\mathcal{A}$ above, the image $\\Psi(\\mathcal{A})$ is Aronhold's degree four $SL_3$ invariant of plane cubics\nand $\\mathcal{A}$ itself is called the (German) symbolic form of that invariant. The classics usually omitted \"$\\Psi(\\cdots)$\" but they \nknew perfectly well what they were doing. Also note that this formalism, in an essentially equivalent form, was invented earlier by Cayley\nas I briefly explained in Section 1.11.4 of my article\n<a href=\"https://arxiv.org/abs/1604.05259\" rel=\"nofollow noreferrer\">\"A Second-Quantized Kolmogorov-Chentsov Theorem\"</a>.\nOne can easily write $\\Psi(\\mathcal{A})$ applied to $F$, in \"physicsy\" tensor contraction notation as\n$$\n\\epsilon_{i_{1} i_{2} i_{3}}\\epsilon_{i_{4} i_{5} i_{6}}\\epsilon_{i_{7} i_{8} i_{9}}\\epsilon_{i_{10} i_{11} i_{12}}\nF_{i_{4} i_{7} i_{10}} F_{i_{1} i_{8} i_{11}} F_{i_{2} i_{5} i_{12}} F_{i_{3} i_{6} i_{9}}\n$$\nwith summation over all twelve indices in the set $\\{1,2,3\\}$. I used the standard physics notation for the Levi-Civita tensor\n$\\epsilon_{i_1 i_2 i_3}$ equal to zero if some indices are equal and otherwise to the sign of the permutation $i_1,i_2,i_3$\nof $1,2,3$. This notation is consistent with the one I used in my previous answers to these MO questions:\n<a href=\"https://mathoverflow.net/questions/121715/invariant-polynomials-for-a-product-of-algebraic-groups\">Q1</a>,\n<a href=\"https://mathoverflow.net/questions/257709/ring-of-invariants-of-operatornamesl-6-acting-on-lambda3-mathbb-c6\">Q2</a>,\nand <a href=\"https://mathoverflow.net/questions/255492/how-to-constructively-combinatorially-prove-schur-weyl-duality\">Q3</a>.\nAt bottom, this is an intrinsically diagrammatic language and if the reader had trouble following what I said so far, it's probably because\nI did not draw any pictures or \"finite-dimensional Feynman diagrams\" (I don't know how to do this on MO). To see these pictures,\nplease have a look at <a href=\"https://arxiv.org/abs/1402.2669\" rel=\"nofollow noreferrer\">this article</a> about an invariant of cubic threefolds or Section 2\nof <a href=\"https://arxiv.org/abs/0904.1734\" rel=\"nofollow noreferrer\">this one</a> mostly on binary forms.\nIn a nutshell, the above contraction can be encoded in a bipartite graph with 4 $\\epsilon$-vertices and 4 $F$-vertices, all of degree three.</p>\n\n<p>Of course the above formalism generalizes with no difficulty (except managing notations) to plethysms of symmetric powers\n$S^p(S^q(\\mathbb{C}^n))$ and similar $\\Phi$, $\\Psi$ maps giving the isomorphism with the relevant space of multisymmetric functions, i.e.,\n$M_{[q,q,\\ldots,q]}^{n}$ (with $q$ repeated $p$ times) or in simpler notation $M_{[q^p]}^{n}$.\nThere is a canonical $GL_n$-equivariant map\n$$\nH_{p,q}:S^p(S^q(\\mathbb{C}^n))\\rightarrow S^q(S^p(\\mathbb{C}^n))\n$$ \nwith \"H\" standing for Hermite, Hadamard or Howe. I will not define it here but it's the first thing that would come to mind\nif asked to produce such a map\n(see <a href=\"https://arxiv.org/abs/1305.7387\" rel=\"nofollow noreferrer\">this review</a> by Landsberg for a detailed discussion in the wider context of geometric complexity theory).\nIt is not hard to see that a homogeneous multisymmetric function in $M_{[p^q]}^{n}\\simeq S^q(S^p(\\mathbb{C}^n))$ is in the image\nof $H_{p,q}$ iff it is a polynomial in the homogeneous elementary multisymmetric functions which belong to $M_{[1^q]}^{n}$.\nIn the context of the Aronhold invariant above, these elementary functions are the coefficients of the polynomial\n$$\nP(x_1,x_2,x_3)=a_x b_x c_x d_x\n$$ \nand similarly in general.</p>\n\n<p>Finally, I can start discussing Will's excellent question. Let $T_{i_1,i_2,i_3}$ denote the tensor of interest with the action of\n$(g,h,k)\\in (SL_n)^3$ given by\n$$\n[(g,h,k)\\cdot T]_{i_1,i_2,i_3}=\\sum_{j_1,j_2,j_3=1}^{n} g_{i_1 j_1} h_{i_2 j_2} k_{i_3 j_3} T_{j_1,j_2,j_3}\\ .\n$$\nIf Will's tensor had an even number $r$ of indices with the action of $(SL_n)^r$ the answer would be trivial.\nTake $n$ copies of $T$, contract all first indices of these copies with one $\\epsilon$,\nthen all second indices with another $\\epsilon$, etc. For $r=3$ this tentative invariant vanishes identically.\nOne can however keep the first indices free, and this gives a map $T\\mapsto \\Gamma(T)=F$ producing an $n$-ary $n$-ic\n$F(x_1,\\ldots,x_n)$ corresponding to the symmetric tensor\n$$\nF_{i_1\\ldots i_n}=\\sum_{j_1,\\ldots,j_n, k_1\\ldots,k_n=1}^{n}\n\\epsilon_{j_1\\ldots j_n} \\epsilon_{k_1\\ldots k_n} T_{i_1,j_1,k_1}\\cdots T_{i_n,j_n,k_n}\\ .\n$$\nFor symmetric $T$'s, such as Will's diagonal tensor $D_{ijk}=\\delta_{ij}\\delta_{ik}$, this is the Hessian covariant.\nJason's idea can thus be rephrased as finding an $SL_n$-invariant $I=I(F)$ of $n$-ary $n$-ics $F$\nand proposing $I(\\Gamma(T))$ as an answer to Will's question. The main issue is to show $I(\\Gamma(D))\\neq 0$.\nNow $\\Gamma(D)$ is the form $F(x_1,\\ldots,x_n)=n!\\ x_1 x_2\\cdots x_n$.</p>\n\n<p>For $n$ even, the smallest degree for an invariant of $n$-ary $n$-ics $F(x_1,\\ldots,x_n)$ is $n$ (unique up to scale). Its symbolic form is\n$$\n\\mathcal{S_n}=(a^{(1)} a^{(2)}\\cdots a^{(n)})^n\n$$\nwhere $a^{(1)}=(a_{1}^{(1)},a_{2}^{(1)}\\ldots,a_{n}^{(1)}),\\ldots,a^{(n)}=(a_{1}^{(n)},a_{2}^{(n)}\\ldots,a_{n}^{(n)})$\nare $n$ series of $n$ variables. Let $I_n=\\Psi(\\mathcal{S_n})$ denote this invariant. One can also write\n$$\nI_n(F)=\\epsilon_{i_{1,1}\\ldots i_{1,n}}\\epsilon_{i_{2,1}\\ldots i_{2,n}}\\cdots\\epsilon_{i_{n,1}\\ldots i_{n,n}}\n\\ F_{i_{1,1}\\ldots i_{n,1}}F_{i_{1,2}\\ldots i_{n,2}}\\cdots\nF_{i_{1,n}\\ldots i_{n,n}}\n$$\nwith summation over indices understood.\nThe evaluation of $I_n(\\Gamma(D))$ is a multiple of the number of row-even minus the number of row-odd $n\\times n$\nLatin squares. Via work of Huang, Rota and Janssen showing that the result is nonzero is equivalent to the (widely open) Alon-Tarsi conjecture.</p>\n\n<p><strong>The following may be new (if not please let me know):</strong>\nFor $n$ odd, the smallest degree for an invariant of $F(x_1,\\ldots,x_n)$ is $n+1$ (unique up to scale). Its symbolic form is\n$$\n\\mathcal{S}=(a^{(2)} a^{(3)}\\cdots a^{(n+1)})(a^{(1)} a^{(3)}\\cdots a^{(n+1)})\\cdots(a^{(1)} a^{(2)}\\cdots a^{(n)})\\ .\n$$\nAgain let me define $I_n=\\Psi(\\mathcal{S_n})$. This is nonzero because the polynomial in the $a^{(i)}$ is not only multihomogeneous\nof multidegree $[n^{n+1}]$ but also multisymmetric. For instance the transposition $a^{(1)}\\leftrightarrow a^{(2)}$ exchanges the first\ntwo bracket factors while the other ones pick up a factor $(-1)$. This gives $(-1)^{n-1}=1$ since $n$ is odd. Other elementary\ntranspositions $a^{(i)}\\leftrightarrow a^{(i+1)}$ can be treated in the same way.\nOf course $I_3$ is Aronhold's degree 4 invariant. I don't know if $I_5$ could be of use for the study of quintic threefolds.\nI boldly (or foolishly?) make the following conjecture which I think is a natural analog of the Alon-Tarsi conjecture in the odd case.</p>\n\n<p><strong>Conjecture:</strong> For $n$ odd, $I_n$ does not vanish on $F(x_1,\\ldots,x_n)=x_1 x_2\\cdots x_n$.</p>\n\n<p>Here is a proof for $n=3$. The Aronhold invariant of degree 4 is the defining equation for the Veronese secant variety\n$\\sigma_{3}(v_3(\\mathbb{P}^2))$.\nHence, it does not vanish on forms with border rank $&gt;3$.\nThe border rank of the form $x_1 x_2\\cdots x_n$ is exactly $2^{n-1}$ as shown by Oeding in <a href=\"https://arxiv.org/abs/1608.02530\" rel=\"nofollow noreferrer\">this article</a>. The case $n=3$ relevant to the present proof was shown earlier by\n<a href=\"https://arxiv.org/abs/0901.0487\" rel=\"nofollow noreferrer\">Landsberg and Teitler</a>. QED.</p>\n\n<p>A final remark regarding the endgame in Jason's argument with a view to producing an explicit invariant of moderate degree: it follows from what I mentioned above that saying that any homogeneous multisymmetric\nfunction in $M_{[p^q]}^{n}\\simeq S^q(S^p(\\mathbb{C}^n))$ can be expressed as a polynomial in the elementary ones is equivalent\nto the Hermite-Hadamard-Howe map $H_{p,q}$ being surjective and this is known for $p$ large with an explicit bound due to Brion (see\nthe review by Landsberg I mentioned). The problem I was discussing is closer to the $p=q$ case related to the Foulkes-Howe conjecture.\nThe latter is known to be false, but not by much! (see <a href=\"https://arxiv.org/abs/1509.03944\" rel=\"nofollow noreferrer\">this article</a>).</p>\n\n<p>Edit: I learned from C. Ikenmeyer that the conjecture about the nonvanishing of $I_n$ for $n$ odd is not new and is stated in Remark 3.26 in\n<a href=\"https://arxiv.org/abs/1511.02927\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1511.02927</a>\nHowever, I noticed recently that this is equivalent to the Alon-Tarsi conjecture for $n$ even. Indeed, the above invariants $I_n$ satisfy the following identity:\n$$\nI_n(F(x_1,\\ldots,x_{n-1})x_n)=(-1)^{\\frac{n}{2}}\\frac{n!}{n^n}\\ I_{n-1}(F)\n$$\nfor all $n$ even and all forms $F$ which only depend on the first $n-1$ variables.</p>\n" }, { "answer_id": 259140, "score": 5, "text": "<p><B>Edit.</B> There was a mistake in the original formulation below of Theorem 2.4, p. 134 of Gelfand-Kapranov-Zelevinsky (now corrected). That theorem only applies after passing to the standard open affines. The upshot is that we know that there exists some integer $d_0&gt;0$ such that the determinant $D_{A,k}^{2d}$ gives a $\\textbf{SL}(A)\\times \\textbf{SL}(B)\\times \\textbf{SL}(C)$-invariant polynomial that is nonzero on Will Sawin's tensor for all $d\\geq d_0$. However, it is not immediately clear what exponent $d_0$ is necessary. An upper bound does follow from the Castelnuovo-Mumford regularity of the symmetric power $(\\mathbb{P}(A))^n//\\mathfrak{S}_n$ as a subvariety of $\\mathbb{P}\\text{Sym}^n_k(A)$, and there are bounds on this by Gotzmann. However, the Gotzmann bound is certainly too large for any effective computations.</p>\n\n<p><B>Original post.</B><br>\nI am just writing up my comments above as an answer. Let $k$ be a commutative, unital ring. Let $A$ be a free $k$-module of rank $n$. Denote by $\\mathbb{P}(A)$ the Proj of the graded $k$-algebra $$S^\\bullet_k(A^\\vee) = \\bigoplus_{d\\geq 0} \\text{Sym}^d_k(A^\\vee).$$<br>\nFor every integer $m\\geq 0$, denote by $(S^\\bullet_k(A^\\vee))^{\\otimes m}$ the $m$-fold tensor product $k$-algebra, $$(S^\\bullet_k(A^\\vee))^{\\otimes m} = S^\\bullet_k(A^\\vee)\\otimes_k \\cdots \\otimes_k S^\\bullet_k(A^\\vee).$$ This is naturally a $\\mathbb{Z}^m$-graded algebra, and it has a natural action of the symmetric group $\\mathfrak{S}_m$. Denote by $(S^\\bullet_k(A^\\vee))^{\\otimes m}_{\\langle 1,\\dots,1 \\rangle}$ the $\\mathbb{Z}$-graded, $\\mathfrak{S}_m$-fixed subalgebra, $$S^\\bullet_k(A^\\vee))^{\\otimes m}_{\\langle 1,\\dots, 1 \\rangle} = \\bigoplus_{d\\geq 0} \\text{Sym}^d_k(A^\\vee)\\otimes_k \\cdots \\otimes_k \\text{Sym}^d_k(A^\\vee).$$\nNow assume that $m!$ is invertible in $k$, so that $\\text{Sym}^m_k(A)^\\vee$ is canonically isomorphic to $\\text{Sym}^m_k(A^\\vee)$ as $k$-vector spaces. Then the transpose of the $m$-multilinear $k$-module homomorphism, $$A^m \\to \\text{Sym}^m_k(A), \\ \\ (a_1,\\dots,a_m)\\mapsto a_1\\cdots a_m,$$ induces a $k$-algebra homomorphism, $$S_k^\\bullet(\\text{Sym}^m_k(A^\\vee)) \\to (S^\\bullet_k(A^\\vee))^{\\otimes m}_{\\langle 1,\\dots,1 \\rangle}.$$ This, in turn, defines a $k$-morphism, $$\\gamma': (\\mathbb{P}(A))^m\\to \\mathbb{P}(\\text{Sym}^m_k(A)),\\ \\ ([a_1],\\dots,[a_m]) \\mapsto [a_1\\cdots a_m]$$ where $(\\mathbb{P}(A))^m$ is shorthand for the $m$-fold self-fiber product $\\mathbb{P}(A)\\times_{\\text{Spec}\\ k}\\cdots \\times_{\\text{Spec}\\ k}\\mathbb{P}(A)$. The morphism $\\gamma'$ is invariant for the natural action of the symmetric group $\\mathfrak{S}_m$ on $(\\mathbb{P}(A))^m$. Thus, $\\gamma'$ factors through the quotient of this group action. Denote by $\\gamma$ the induced $k$-morphism, $$\\gamma: (\\mathbb{P}(A))^m//\\mathfrak{S}_m \\to \\mathbb{P}(\\text{Sym}^m_k(A)).$$</p>\n\n<p>It is a classical fact that $\\gamma$ is a closed immersion. This is explicitly stated and proved in Proposition 2.1 and Theorem 2.2, Chapter 4, pp. 132-133 of Gelfand-Kapranov-Zelevinsky. In fact, Gelfand-Kapranov-Zelevinsky prove much more: for every nonzero $x_i\\in A^\\vee$, denoting by $\\gamma_m(x_i)\\in \\text{Sym}^m_k(A^\\vee)$ the associated $m^{\\text{th}}$ power, then the induced $k$-algebra homomorphism of affine opens, $$S_k^\\bullet(\\text{Sym}^m_k(A^\\vee))[1/\\gamma_m(x_i)] \\to [(S_k^\\bullet(A^\\vee))^{\\otimes m}_{\\langle 1,\\dots,1 \\rangle}]^{\\mathfrak{S}_m}[1/(x_{1,i}\\cdots x_{m,i})],$$ is a surjective $k$-algebra homomorphism. This is essentially Theorem 2.4 on p. 134. From this it follows that for every $\\mathfrak{S}_m$-invariant, homogeneous element $D$ of degree $r&gt;0$ in $(S_k^\\bullet(A^\\vee))^{\\otimes m}_{(r,\\dots,r)}$, there exists an integer $e_0$ such that for every $e\\geq e_0$, $D$ times the image of $\\text{Sym}^e_k(\\text{Sym}^m_k(A^\\vee))$ is in the image of $\\text{Sym}^{e+r}_k(\\text{Sym}^m_k(A^\\vee))$. Thus there exists an integer $d&gt;0$ such that $D^d$ is in the image of $\\text{Sym}^{dr}_k(\\text{Sym}^m_k(A^\\vee))$.</p>\n\n<p>In particular, for $m=n$ (assuming that $n!$ is invertible in $k$), the determinant map, $$\\text{det}_{A,k}: A^n \\to \\bigwedge^n_k(A),$$ defines an element $D_{A,k} \\in (S_k^\\bullet(A^\\vee))^{\\otimes n}_{(1,\\dots,1)}$. This element is <B>not</B> $\\mathfrak{S}_m$-invariant. However, the square, $D_{A,k}^2$ is $\\mathfrak{S}_m$-invariant. Thus, $D_{A,k}^{2d}$ is in the image $I_{2d}$ of $\\text{Sym}_k^{2d}(\\text{Sym}_k^n(A^\\vee))$ for some integer $d&gt;0$.</p>\n\n<p>Finally, if we assume that $k$ is a field of characteristic $0$, then $\\textbf{SL}(A)$ is linearly reductive. Thus, the induced $k$-linear map, $$\\text{Sym}^{2d}_k(\\text{Sym}^m_k(A^\\vee))^{\\textbf{SL}(A)} \\to I_{2d}^{\\textbf{SL}(A)},$$\nis surjective. In particular, the $\\textbf{SL}(A)$-invariant element $D_{A,k}^{2d}$ is in the image of $\\left( \\text{Sym}_k^{2d}(\\text{Sym}_k^n(A^\\vee)) \\right)^{\\textbf{SL}(A)}$. So there exists some $\\textbf{SL}(A)$-invariant element $\\Delta_{A,k}$ mappsing to $D_{A,k}^{2d}$.</p>\n\n<p>Now put this together with the tensor operation from the comments. For $k$-vector spaces $A$, $B$, and $C$ of dimension $n$, for every $k$-linear map, $$t:A^\\vee \\to \\text{Hom}_k(B,C),$$ there is an associated element $$\\widetilde{t} \\in \\text{Sym}^n(A) \\otimes_k \\text{Hom}_k(\\bigwedge^n_k(B),\\bigwedge^n_k(C)),$$ giving the polynomial map, $$ A^\\vee\\to \\text{Hom}_k(\\bigwedge^n_k(B),\\bigwedge^n_k(C)), \\ \\ a \\mapsto \\text{det}(t(a)).$$ The associated map $$\\text{Hom}_k(A^\\vee,\\text{Hom}_k(B,C))\\to \\text{Sym}^n(A)\\otimes_k \\text{Hom}_k(\\bigwedge^n_k(B),\\bigwedge^n_k(C)), \\ \\ t\\mapsto \\widetilde{t},$$ is a degree $n$, homogeneous $k$-polynomial map that is equivariant for the natural actions of $\\textbf{SL}(A)\\times \\textbf{SL}(B)\\times \\textbf{SL}(C)$. Via the duality of $\\text{Sym}^n(A)$ and $\\text{Sym}^n(A^\\vee)$, the polynomial $\\Delta_{A,k}$ determines a $\\textbf{SL}(A)\\times \\textbf{SL}(B)\\times \\textbf{SL}(C)$-invariant homogeneous polynomial of degree $2dn$ on $\\text{Hom}_k(A^\\vee,\\text{Hom}_k(B,C))$. </p>\n\n<p>For the specific tensor $t$ that you wrote down, with respect to appropriate bases for $A$, $B$, and $C$, for every $(a_1,\\dots,a_n)\\in A$, $t(a_1,\\dots,a_n)$ is the linear transformation whose matrix representative is the diagonal $n\\times n$ matrix with entries $(a_1,\\dots,a_n)$. Thus, $\\widetilde{t}$ equals $(a_1\\cdots a_n)\\otimes (b_1\\wedge \\dots \\wedge b_n)\\otimes (c_1\\wedge \\dots c_n)$. In particular, since $a_1\\cdots a_n$ is in the image of $\\gamma$, $\\Delta_{A,k}$ restricts on $\\widetilde{t}$ as a power of the determinant. Since $a_1\\cdots a_n$ is a product of $n$ linearly independent linear polynomials, the determinant is nonzero. Thus, $\\Delta_{A,k}$ is a $\\textbf{SL}(A)\\times \\textbf{SL}(B)\\times \\textbf{SL}(C)$-invariant, homogeneous polynomial of degree $2n$ that is nonzero on your tensor. </p>\n" } ]
Question: Explicit invariant of tensors nonvanishing on the diagonal Answer 1: <p>Let me start with some remarks about the classical symbolic method (without which one cannot understand 19th century invariant theory) and multisymmetric functions. I will use an example first. Take four series of three variables $a=(a_1,a_2,a_3)$, $b=(b_1,b_2,b_3)$, $c=(c_1,c_2,c_3)$ and $d=(d_1,d_2,d_3)$. Now define the polynomial $\mathcal{A}$ in these 12 indeterminates given by $$ \mathcal{A}=(bcd)(acd)(abd)(abc) $$ where we used a shorthand notation for determinants $$ (bcd)=\left| \begin{array}{ccc} b_1 &amp; b_2 &amp; b_3 \\ c_1 &amp; c_2 &amp; c_3 \\ d_1 &amp; d_2 &amp; d_3 \\ \end{array} \right| $$ and similarly for the other "bracket factors" $(acd)$, etc. This is an example of multisymmetric function in the "letters" $a,b,c,d$ which is multihomogeneous of multidegree $[3,3,3,3]$. Let me write $M_{[3,3,3,3]}^3$ for the space of such multisymmetric functions. The subscript is the multidegree while the superscript refers to the number of variables in each series. The symmetric group acting is $\mathfrak{S}_4$ which permutes these series. If each series had one variable in it instead of three we would be in the familiar realm of the theory of ordinary symmetric functions. The analogous theory of multisymmetric functions is much more complicated and has been studied classically by Euler, Lagrange, Poisson, Schl&auml;fli, Brill, Gordan, Junker and more recently by Angeniol, Dalbec, Briand, Vaccarino, Domokos and others. A nice yet old reference on this is the book on elimination theory by Faa di Bruno I mentioned in <a href="https://mathoverflow.net/questions/117415/old-books-still-used/117417#117417">this MO answer</a>.</p> <p>Consider now a generic ternary cubic form $F=F(x_1,x_2,x_3)$. One can write it as $$ F(x)=\sum_{i_1,i_2,i_3=1}^3 F_{i_1,i_2,i_3} x_{i_1} x_{i_2} x_{i_3} $$ where $(F_{i_1,i_2,i_3})$ is a symmetric tensor (by which I mean a "matrix" rather than a Bourbaki tensor). This form lives in the symmetric power $S^3(\mathbb{C}^3)$ but I will be (intentionally) sloppy and write $S^3$ instead. The space of degree 4 homogeneous polynomials in the coefficients of $F$ will be denoted by $S^4(S^3)$.</p> <p>Let me define a linear map $\Phi:S^4(S^3)\rightarrow M_{[3,3,3,3]}^3$ as follows. Given a degree four homogeneous polynomial $C(F)$ first construct its full polarization $$ \widetilde{C}(F_1,F_2,F_3,F_4)=\frac{1}{4!} \frac{\partial^4}{\partial t_1\partial t_2\partial t_3\partial t_4} C(t_1F_1+t_2 F_2+t_3 F_3+t_4 F_4) $$ which satisfies $\widetilde{C}(F,F,F,F)=C(F)$. Now set $F_1(x)=a_{x}^3$ where $a_x$ is the classical notation for the linear form $a_1x_1+a_2x_2+a_3x_3$ and similarly specialize to the cubes of linear forms $F_2(x)=b_{x}^3$, $F_3(x)=c_{x}^3$ and $F_4(x)=d_{x}^3$. Finally, define $\Phi(C)=\widetilde{C}(F_1,F_2,F_3,F_4)$.</p> <p>I will also define another linear map $\Psi: M_{[3,3,3,3]}^3\rightarrow S^4(S^3)$. Let $\mathcal{S}=\mathcal{S}(a,b,c,d)$ be a multisymmetric function in $M_{[3,3,3,3]}^3$. We define the differential operator $$ \mathcal{D}=\frac{1}{3!^4} F\left(\frac{\partial}{\partial a_1},\frac{\partial}{\partial a_2},\frac{\partial}{\partial a_3}\right) F\left(\frac{\partial}{\partial b_1},\frac{\partial}{\partial b_2},\frac{\partial}{\partial b_3}\right) $$ $$ F\left(\frac{\partial}{\partial c_1},\frac{\partial}{\partial c_2},\frac{\partial}{\partial c_3}\right) F\left(\frac{\partial}{\partial d_1},\frac{\partial}{\partial d_2},\frac{\partial}{\partial d_3}\right) $$ and let $\Psi(\mathcal{S})=\mathcal{D}\ \mathcal{S}$, i.e., the result of applying the differential operator $\mathcal{D}$ to the polynomial $\mathcal{S}$.</p> <p>It is not hard to see that $\Phi$ and $\Psi$ are inverses of one another. In the particular case of $\mathcal{A}$ above, the image $\Psi(\mathcal{A})$ is Aronhold's degree four $SL_3$ invariant of plane cubics and $\mathcal{A}$ itself is called the (German) symbolic form of that invariant. The classics usually omitted "$\Psi(\cdots)$" but they knew perfectly well what they were doing. Also note that this formalism, in an essentially equivalent form, was invented earlier by Cayley as I briefly explained in Section 1.11.4 of my article <a href="https://arxiv.org/abs/1604.05259" rel="nofollow noreferrer">"A Second-Quantized Kolmogorov-Chentsov Theorem"</a>. One can easily write $\Psi(\mathcal{A})$ applied to $F$, in "physicsy" tensor contraction notation as $$ \epsilon_{i_{1} i_{2} i_{3}}\epsilon_{i_{4} i_{5} i_{6}}\epsilon_{i_{7} i_{8} i_{9}}\epsilon_{i_{10} i_{11} i_{12}} F_{i_{4} i_{7} i_{10}} F_{i_{1} i_{8} i_{11}} F_{i_{2} i_{5} i_{12}} F_{i_{3} i_{6} i_{9}} $$ with summation over all twelve indices in the set $\{1,2,3\}$. I used the standard physics notation for the Levi-Civita tensor $\epsilon_{i_1 i_2 i_3}$ equal to zero if some indices are equal and otherwise to the sign of the permutation $i_1,i_2,i_3$ of $1,2,3$. This notation is consistent with the one I used in my previous answers to these MO questions: <a href="https://mathoverflow.net/questions/121715/invariant-polynomials-for-a-product-of-algebraic-groups">Q1</a>, <a href="https://mathoverflow.net/questions/257709/ring-of-invariants-of-operatornamesl-6-acting-on-lambda3-mathbb-c6">Q2</a>, and <a href="https://mathoverflow.net/questions/255492/how-to-constructively-combinatorially-prove-schur-weyl-duality">Q3</a>. At bottom, this is an intrinsically diagrammatic language and if the reader had trouble following what I said so far, it's probably because I did not draw any pictures or "finite-dimensional Feynman diagrams" (I don't know how to do this on MO). To see these pictures, please have a look at <a href="https://arxiv.org/abs/1402.2669" rel="nofollow noreferrer">this article</a> about an invariant of cubic threefolds or Section 2 of <a href="https://arxiv.org/abs/0904.1734" rel="nofollow noreferrer">this one</a> mostly on binary forms. In a nutshell, the above contraction can be encoded in a bipartite graph with 4 $\epsilon$-vertices and 4 $F$-vertices, all of degree three.</p> <p>Of course the above formalism generalizes with no difficulty (except managing notations) to plethysms of symmetric powers $S^p(S^q(\mathbb{C}^n))$ and similar $\Phi$, $\Psi$ maps giving the isomorphism with the relevant space of multisymmetric functions, i.e., $M_{[q,q,\ldots,q]}^{n}$ (with $q$ repeated $p$ times) or in simpler notation $M_{[q^p]}^{n}$. There is a canonical $GL_n$-equivariant map $$ H_{p,q}:S^p(S^q(\mathbb{C}^n))\rightarrow S^q(S^p(\mathbb{C}^n)) $$ with "H" standing for Hermite, Hadamard or Howe. I will not define it here but it's the first thing that would come to mind if asked to produce such a map (see <a href="https://arxiv.org/abs/1305.7387" rel="nofollow noreferrer">this review</a> by Landsberg for a detailed discussion in the wider context of geometric complexity theory). It is not hard to see that a homogeneous multisymmetric function in $M_{[p^q]}^{n}\simeq S^q(S^p(\mathbb{C}^n))$ is in the image of $H_{p,q}$ iff it is a polynomial in the homogeneous elementary multisymmetric functions which belong to $M_{[1^q]}^{n}$. In the context of the Aronhold invariant above, these elementary functions are the coefficients of the polynomial $$ P(x_1,x_2,x_3)=a_x b_x c_x d_x $$ and similarly in general.</p> <p>Finally, I can start discussing Will's excellent question. Let $T_{i_1,i_2,i_3}$ denote the tensor of interest with the action of $(g,h,k)\in (SL_n)^3$ given by $$ [(g,h,k)\cdot T]_{i_1,i_2,i_3}=\sum_{j_1,j_2,j_3=1}^{n} g_{i_1 j_1} h_{i_2 j_2} k_{i_3 j_3} T_{j_1,j_2,j_3}\ . $$ If Will's tensor had an even number $r$ of indices with the action of $(SL_n)^r$ the answer would be trivial. Take $n$ copies of $T$, contract all first indices of these copies with one $\epsilon$, then all second indices with another $\epsilon$, etc. For $r=3$ this tentative invariant vanishes identically. One can however keep the first indices free, and this gives a map $T\mapsto \Gamma(T)=F$ producing an $n$-ary $n$-ic $F(x_1,\ldots,x_n)$ corresponding to the symmetric tensor $$ F_{i_1\ldots i_n}=\sum_{j_1,\ldots,j_n, k_1\ldots,k_n=1}^{n} \epsilon_{j_1\ldots j_n} \epsilon_{k_1\ldots k_n} T_{i_1,j_1,k_1}\cdots T_{i_n,j_n,k_n}\ . $$ For symmetric $T$'s, such as Will's diagonal tensor $D_{ijk}=\delta_{ij}\delta_{ik}$, this is the Hessian covariant. Jason's idea can thus be rephrased as finding an $SL_n$-invariant $I=I(F)$ of $n$-ary $n$-ics $F$ and proposing $I(\Gamma(T))$ as an answer to Will's question. The main issue is to show $I(\Gamma(D))\neq 0$. Now $\Gamma(D)$ is the form $F(x_1,\ldots,x_n)=n!\ x_1 x_2\cdots x_n$.</p> <p>For $n$ even, the smallest degree for an invariant of $n$-ary $n$-ics $F(x_1,\ldots,x_n)$ is $n$ (unique up to scale). Its symbolic form is $$ \mathcal{S_n}=(a^{(1)} a^{(2)}\cdots a^{(n)})^n $$ where $a^{(1)}=(a_{1}^{(1)},a_{2}^{(1)}\ldots,a_{n}^{(1)}),\ldots,a^{(n)}=(a_{1}^{(n)},a_{2}^{(n)}\ldots,a_{n}^{(n)})$ are $n$ series of $n$ variables. Let $I_n=\Psi(\mathcal{S_n})$ denote this invariant. One can also write $$ I_n(F)=\epsilon_{i_{1,1}\ldots i_{1,n}}\epsilon_{i_{2,1}\ldots i_{2,n}}\cdots\epsilon_{i_{n,1}\ldots i_{n,n}} \ F_{i_{1,1}\ldots i_{n,1}}F_{i_{1,2}\ldots i_{n,2}}\cdots F_{i_{1,n}\ldots i_{n,n}} $$ with summation over indices understood. The evaluation of $I_n(\Gamma(D))$ is a multiple of the number of row-even minus the number of row-odd $n\times n$ Latin squares. Via work of Huang, Rota and Janssen showing that the result is nonzero is equivalent to the (widely open) Alon-Tarsi conjecture.</p> <p><strong>The following may be new (if not please let me know):</strong> For $n$ odd, the smallest degree for an invariant of $F(x_1,\ldots,x_n)$ is $n+1$ (unique up to scale). Its symbolic form is $$ \mathcal{S}=(a^{(2)} a^{(3)}\cdots a^{(n+1)})(a^{(1)} a^{(3)}\cdots a^{(n+1)})\cdots(a^{(1)} a^{(2)}\cdots a^{(n)})\ . $$ Again let me define $I_n=\Psi(\mathcal{S_n})$. This is nonzero because the polynomial in the $a^{(i)}$ is not only multihomogeneous of multidegree $[n^{n+1}]$ but also multisymmetric. For instance the transposition $a^{(1)}\leftrightarrow a^{(2)}$ exchanges the first two bracket factors while the other ones pick up a factor $(-1)$. This gives $(-1)^{n-1}=1$ since $n$ is odd. Other elementary transpositions $a^{(i)}\leftrightarrow a^{(i+1)}$ can be treated in the same way. Of course $I_3$ is Aronhold's degree 4 invariant. I don't know if $I_5$ could be of use for the study of quintic threefolds. I boldly (or foolishly?) make the following conjecture which I think is a natural analog of the Alon-Tarsi conjecture in the odd case.</p> <p><strong>Conjecture:</strong> For $n$ odd, $I_n$ does not vanish on $F(x_1,\ldots,x_n)=x_1 x_2\cdots x_n$.</p> <p>Here is a proof for $n=3$. The Aronhold invariant of degree 4 is the defining equation for the Veronese secant variety $\sigma_{3}(v_3(\mathbb{P}^2))$. Hence, it does not vanish on forms with border rank $&gt;3$. The border rank of the form $x_1 x_2\cdots x_n$ is exactly $2^{n-1}$ as shown by Oeding in <a href="https://arxiv.org/abs/1608.02530" rel="nofollow noreferrer">this article</a>. The case $n=3$ relevant to the present proof was shown earlier by <a href="https://arxiv.org/abs/0901.0487" rel="nofollow noreferrer">Landsberg and Teitler</a>. QED.</p> <p>A final remark regarding the endgame in Jason's argument with a view to producing an explicit invariant of moderate degree: it follows from what I mentioned above that saying that any homogeneous multisymmetric function in $M_{[p^q]}^{n}\simeq S^q(S^p(\mathbb{C}^n))$ can be expressed as a polynomial in the elementary ones is equivalent to the Hermite-Hadamard-Howe map $H_{p,q}$ being surjective and this is known for $p$ large with an explicit bound due to Brion (see the review by Landsberg I mentioned). The problem I was discussing is closer to the $p=q$ case related to the Foulkes-Howe conjecture. The latter is known to be false, but not by much! (see <a href="https://arxiv.org/abs/1509.03944" rel="nofollow noreferrer">this article</a>).</p> <p>Edit: I learned from C. Ikenmeyer that the conjecture about the nonvanishing of $I_n$ for $n$ odd is not new and is stated in Remark 3.26 in <a href="https://arxiv.org/abs/1511.02927" rel="nofollow noreferrer">https://arxiv.org/abs/1511.02927</a> However, I noticed recently that this is equivalent to the Alon-Tarsi conjecture for $n$ even. Indeed, the above invariants $I_n$ satisfy the following identity: $$ I_n(F(x_1,\ldots,x_{n-1})x_n)=(-1)^{\frac{n}{2}}\frac{n!}{n^n}\ I_{n-1}(F) $$ for all $n$ even and all forms $F$ which only depend on the first $n-1$ variables.</p> Answer 2: <p><B>Edit.</B> There was a mistake in the original formulation below of Theorem 2.4, p. 134 of Gelfand-Kapranov-Zelevinsky (now corrected). That theorem only applies after passing to the standard open affines. The upshot is that we know that there exists some integer $d_0&gt;0$ such that the determinant $D_{A,k}^{2d}$ gives a $\textbf{SL}(A)\times \textbf{SL}(B)\times \textbf{SL}(C)$-invariant polynomial that is nonzero on Will Sawin's tensor for all $d\geq d_0$. However, it is not immediately clear what exponent $d_0$ is necessary. An upper bound does follow from the Castelnuovo-Mumford regularity of the symmetric power $(\mathbb{P}(A))^n//\mathfrak{S}_n$ as a subvariety of $\mathbb{P}\text{Sym}^n_k(A)$, and there are bounds on this by Gotzmann. However, the Gotzmann bound is certainly too large for any effective computations.</p> <p><B>Original post.</B><br> I am just writing up my comments above as an answer. Let $k$ be a commutative, unital ring. Let $A$ be a free $k$-module of rank $n$. Denote by $\mathbb{P}(A)$ the Proj of the graded $k$-algebra $$S^\bullet_k(A^\vee) = \bigoplus_{d\geq 0} \text{Sym}^d_k(A^\vee).$$<br> For every integer $m\geq 0$, denote by $(S^\bullet_k(A^\vee))^{\otimes m}$ the $m$-fold tensor product $k$-algebra, $$(S^\bullet_k(A^\vee))^{\otimes m} = S^\bullet_k(A^\vee)\otimes_k \cdots \otimes_k S^\bullet_k(A^\vee).$$ This is naturally a $\mathbb{Z}^m$-graded algebra, and it has a natural action of the symmetric group $\mathfrak{S}_m$. Denote by $(S^\bullet_k(A^\vee))^{\otimes m}_{\langle 1,\dots,1 \rangle}$ the $\mathbb{Z}$-graded, $\mathfrak{S}_m$-fixed subalgebra, $$S^\bullet_k(A^\vee))^{\otimes m}_{\langle 1,\dots, 1 \rangle} = \bigoplus_{d\geq 0} \text{Sym}^d_k(A^\vee)\otimes_k \cdots \otimes_k \text{Sym}^d_k(A^\vee).$$ Now assume that $m!$ is invertible in $k$, so that $\text{Sym}^m_k(A)^\vee$ is canonically isomorphic to $\text{Sym}^m_k(A^\vee)$ as $k$-vector spaces. Then the transpose of the $m$-multilinear $k$-module homomorphism, $$A^m \to \text{Sym}^m_k(A), \ \ (a_1,\dots,a_m)\mapsto a_1\cdots a_m,$$ induces a $k$-algebra homomorphism, $$S_k^\bullet(\text{Sym}^m_k(A^\vee)) \to (S^\bullet_k(A^\vee))^{\otimes m}_{\langle 1,\dots,1 \rangle}.$$ This, in turn, defines a $k$-morphism, $$\gamma': (\mathbb{P}(A))^m\to \mathbb{P}(\text{Sym}^m_k(A)),\ \ ([a_1],\dots,[a_m]) \mapsto [a_1\cdots a_m]$$ where $(\mathbb{P}(A))^m$ is shorthand for the $m$-fold self-fiber product $\mathbb{P}(A)\times_{\text{Spec}\ k}\cdots \times_{\text{Spec}\ k}\mathbb{P}(A)$. The morphism $\gamma'$ is invariant for the natural action of the symmetric group $\mathfrak{S}_m$ on $(\mathbb{P}(A))^m$. Thus, $\gamma'$ factors through the quotient of this group action. Denote by $\gamma$ the induced $k$-morphism, $$\gamma: (\mathbb{P}(A))^m//\mathfrak{S}_m \to \mathbb{P}(\text{Sym}^m_k(A)).$$</p> <p>It is a classical fact that $\gamma$ is a closed immersion. This is explicitly stated and proved in Proposition 2.1 and Theorem 2.2, Chapter 4, pp. 132-133 of Gelfand-Kapranov-Zelevinsky. In fact, Gelfand-Kapranov-Zelevinsky prove much more: for every nonzero $x_i\in A^\vee$, denoting by $\gamma_m(x_i)\in \text{Sym}^m_k(A^\vee)$ the associated $m^{\text{th}}$ power, then the induced $k$-algebra homomorphism of affine opens, $$S_k^\bullet(\text{Sym}^m_k(A^\vee))[1/\gamma_m(x_i)] \to [(S_k^\bullet(A^\vee))^{\otimes m}_{\langle 1,\dots,1 \rangle}]^{\mathfrak{S}_m}[1/(x_{1,i}\cdots x_{m,i})],$$ is a surjective $k$-algebra homomorphism. This is essentially Theorem 2.4 on p. 134. From this it follows that for every $\mathfrak{S}_m$-invariant, homogeneous element $D$ of degree $r&gt;0$ in $(S_k^\bullet(A^\vee))^{\otimes m}_{(r,\dots,r)}$, there exists an integer $e_0$ such that for every $e\geq e_0$, $D$ times the image of $\text{Sym}^e_k(\text{Sym}^m_k(A^\vee))$ is in the image of $\text{Sym}^{e+r}_k(\text{Sym}^m_k(A^\vee))$. Thus there exists an integer $d&gt;0$ such that $D^d$ is in the image of $\text{Sym}^{dr}_k(\text{Sym}^m_k(A^\vee))$.</p> <p>In particular, for $m=n$ (assuming that $n!$ is invertible in $k$), the determinant map, $$\text{det}_{A,k}: A^n \to \bigwedge^n_k(A),$$ defines an element $D_{A,k} \in (S_k^\bullet(A^\vee))^{\otimes n}_{(1,\dots,1)}$. This element is <B>not</B> $\mathfrak{S}_m$-invariant. However, the square, $D_{A,k}^2$ is $\mathfrak{S}_m$-invariant. Thus, $D_{A,k}^{2d}$ is in the image $I_{2d}$ of $\text{Sym}_k^{2d}(\text{Sym}_k^n(A^\vee))$ for some integer $d&gt;0$.</p> <p>Finally, if we assume that $k$ is a field of characteristic $0$, then $\textbf{SL}(A)$ is linearly reductive. Thus, the induced $k$-linear map, $$\text{Sym}^{2d}_k(\text{Sym}^m_k(A^\vee))^{\textbf{SL}(A)} \to I_{2d}^{\textbf{SL}(A)},$$ is surjective. In particular, the $\textbf{SL}(A)$-invariant element $D_{A,k}^{2d}$ is in the image of $\left( \text{Sym}_k^{2d}(\text{Sym}_k^n(A^\vee)) \right)^{\textbf{SL}(A)}$. So there exists some $\textbf{SL}(A)$-invariant element $\Delta_{A,k}$ mappsing to $D_{A,k}^{2d}$.</p> <p>Now put this together with the tensor operation from the comments. For $k$-vector spaces $A$, $B$, and $C$ of dimension $n$, for every $k$-linear map, $$t:A^\vee \to \text{Hom}_k(B,C),$$ there is an associated element $$\widetilde{t} \in \text{Sym}^n(A) \otimes_k \text{Hom}_k(\bigwedge^n_k(B),\bigwedge^n_k(C)),$$ giving the polynomial map, $$ A^\vee\to \text{Hom}_k(\bigwedge^n_k(B),\bigwedge^n_k(C)), \ \ a \mapsto \text{det}(t(a)).$$ The associated map $$\text{Hom}_k(A^\vee,\text{Hom}_k(B,C))\to \text{Sym}^n(A)\otimes_k \text{Hom}_k(\bigwedge^n_k(B),\bigwedge^n_k(C)), \ \ t\mapsto \widetilde{t},$$ is a degree $n$, homogeneous $k$-polynomial map that is equivariant for the natural actions of $\textbf{SL}(A)\times \textbf{SL}(B)\times \textbf{SL}(C)$. Via the duality of $\text{Sym}^n(A)$ and $\text{Sym}^n(A^\vee)$, the polynomial $\Delta_{A,k}$ determines a $\textbf{SL}(A)\times \textbf{SL}(B)\times \textbf{SL}(C)$-invariant homogeneous polynomial of degree $2dn$ on $\text{Hom}_k(A^\vee,\text{Hom}_k(B,C))$. </p> <p>For the specific tensor $t$ that you wrote down, with respect to appropriate bases for $A$, $B$, and $C$, for every $(a_1,\dots,a_n)\in A$, $t(a_1,\dots,a_n)$ is the linear transformation whose matrix representative is the diagonal $n\times n$ matrix with entries $(a_1,\dots,a_n)$. Thus, $\widetilde{t}$ equals $(a_1\cdots a_n)\otimes (b_1\wedge \dots \wedge b_n)\otimes (c_1\wedge \dots c_n)$. In particular, since $a_1\cdots a_n$ is in the image of $\gamma$, $\Delta_{A,k}$ restricts on $\widetilde{t}$ as a power of the determinant. Since $a_1\cdots a_n$ is a product of $n$ linearly independent linear polynomials, the determinant is nonzero. Thus, $\Delta_{A,k}$ is a $\textbf{SL}(A)\times \textbf{SL}(B)\times \textbf{SL}(C)$-invariant, homogeneous polynomial of degree $2n$ that is nonzero on your tensor. </p> Answer 3: <p>Let me start with some remarks about the classical symbolic method (without which one cannot understand 19th century invariant theory) and multisymmetric functions. I will use an example first. Take four series of three variables $a=(a_1,a_2,a_3)$, $b=(b_1,b_2,b_3)$, $c=(c_1,c_2,c_3)$ and $d=(d_1,d_2,d_3)$. Now define the polynomial $\mathcal{A}$ in these 12 indeterminates given by $$ \mathcal{A}=(bcd)(acd)(abd)(abc) $$ where we used a shorthand notation for determinants $$ (bcd)=\left| \begin{array}{ccc} b_1 &amp; b_2 &amp; b_3 \\ c_1 &amp; c_2 &amp; c_3 \\ d_1 &amp; d_2 &amp; d_3 \\ \end{array} \right| $$ and similarly for the other "bracket factors" $(acd)$, etc. This is an example of multisymmetric function in the "letters" $a,b,c,d$ which is multihomogeneous of multidegree $[3,3,3,3]$. Let me write $M_{[3,3,3,3]}^3$ for the space of such multisymmetric functions. The subscript is the multidegree while the superscript refers to the number of variables in each series. The symmetric group acting is $\mathfrak{S}_4$ which permutes these series. If each series had one variable in it instead of three we would be in the familiar realm of the theory of ordinary symmetric functions. The analogous theory of multisymmetric functions is much more complicated and has been studied classically by Euler, Lagrange, Poisson, Schl&auml;fli, Brill, Gordan, Junker and more recently by Angeniol, Dalbec, Briand, Vaccarino, Domokos and others. A nice yet old reference on this is the book on elimination theory by Faa di Bruno I mentioned in <a href="https://mathoverflow.net/questions/117415/old-books-still-used/117417#117417">this MO answer</a>.</p> <p>Consider now a generic ternary cubic form $F=F(x_1,x_2,x_3)$. One can write it as $$ F(x)=\sum_{i_1,i_2,i_3=1}^3 F_{i_1,i_2,i_3} x_{i_1} x_{i_2} x_{i_3} $$ where $(F_{i_1,i_2,i_3})$ is a symmetric tensor (by which I mean a "matrix" rather than a Bourbaki tensor). This form lives in the symmetric power $S^3(\mathbb{C}^3)$ but I will be (intentionally) sloppy and write $S^3$ instead. The space of degree 4 homogeneous polynomials in the coefficients of $F$ will be denoted by $S^4(S^3)$.</p> <p>Let me define a linear map $\Phi:S^4(S^3)\rightarrow M_{[3,3,3,3]}^3$ as follows. Given a degree four homogeneous polynomial $C(F)$ first construct its full polarization $$ \widetilde{C}(F_1,F_2,F_3,F_4)=\frac{1}{4!} \frac{\partial^4}{\partial t_1\partial t_2\partial t_3\partial t_4} C(t_1F_1+t_2 F_2+t_3 F_3+t_4 F_4) $$ which satisfies $\widetilde{C}(F,F,F,F)=C(F)$. Now set $F_1(x)=a_{x}^3$ where $a_x$ is the classical notation for the linear form $a_1x_1+a_2x_2+a_3x_3$ and similarly specialize to the cubes of linear forms $F_2(x)=b_{x}^3$, $F_3(x)=c_{x}^3$ and $F_4(x)=d_{x}^3$. Finally, define $\Phi(C)=\widetilde{C}(F_1,F_2,F_3,F_4)$.</p> <p>I will also define another linear map $\Psi: M_{[3,3,3,3]}^3\rightarrow S^4(S^3)$. Let $\mathcal{S}=\mathcal{S}(a,b,c,d)$ be a multisymmetric function in $M_{[3,3,3,3]}^3$. We define the differential operator $$ \mathcal{D}=\frac{1}{3!^4} F\left(\frac{\partial}{\partial a_1},\frac{\partial}{\partial a_2},\frac{\partial}{\partial a_3}\right) F\left(\frac{\partial}{\partial b_1},\frac{\partial}{\partial b_2},\frac{\partial}{\partial b_3}\right) $$ $$ F\left(\frac{\partial}{\partial c_1},\frac{\partial}{\partial c_2},\frac{\partial}{\partial c_3}\right) F\left(\frac{\partial}{\partial d_1},\frac{\partial}{\partial d_2},\frac{\partial}{\partial d_3}\right) $$ and let $\Psi(\mathcal{S})=\mathcal{D}\ \mathcal{S}$, i.e., the result of applying the differential operator $\mathcal{D}$ to the polynomial $\mathcal{S}$.</p> <p>It is not hard to see that $\Phi$ and $\Psi$ are inverses of one another. In the particular case of $\mathcal{A}$ above, the image $\Psi(\mathcal{A})$ is Aronhold's degree four $SL_3$ invariant of plane cubics and $\mathcal{A}$ itself is called the (German) symbolic form of that invariant. The classics usually omitted "$\Psi(\cdots)$" but they knew perfectly well what they were doing. Also note that this formalism, in an essentially equivalent form, was invented earlier by Cayley as I briefly explained in Section 1.11.4 of my article <a href="https://arxiv.org/abs/1604.05259" rel="nofollow noreferrer">"A Second-Quantized Kolmogorov-Chentsov Theorem"</a>. One can easily write $\Psi(\mathcal{A})$ applied to $F$, in "physicsy" tensor contraction notation as $$ \epsilon_{i_{1} i_{2} i_{3}}\epsilon_{i_{4} i_{5} i_{6}}\epsilon_{i_{7} i_{8} i_{9}}\epsilon_{i_{10} i_{11} i_{12}} F_{i_{4} i_{7} i_{10}} F_{i_{1} i_{8} i_{11}} F_{i_{2} i_{5} i_{12}} F_{i_{3} i_{6} i_{9}} $$ with summation over all twelve indices in the set $\{1,2,3\}$. I used the standard physics notation for the Levi-Civita tensor $\epsilon_{i_1 i_2 i_3}$ equal to zero if some indices are equal and otherwise to the sign of the permutation $i_1,i_2,i_3$ of $1,2,3$. This notation is consistent with the one I used in my previous answers to these MO questions: <a href="https://mathoverflow.net/questions/121715/invariant-polynomials-for-a-product-of-algebraic-groups">Q1</a>, <a href="https://mathoverflow.net/questions/257709/ring-of-invariants-of-operatornamesl-6-acting-on-lambda3-mathbb-c6">Q2</a>, and <a href="https://mathoverflow.net/questions/255492/how-to-constructively-combinatorially-prove-schur-weyl-duality">Q3</a>. At bottom, this is an intrinsically diagrammatic language and if the reader had trouble following what I said so far, it's probably because I did not draw any pictures or "finite-dimensional Feynman diagrams" (I don't know how to do this on MO). To see these pictures, please have a look at <a href="https://arxiv.org/abs/1402.2669" rel="nofollow noreferrer">this article</a> about an invariant of cubic threefolds or Section 2 of <a href="https://arxiv.org/abs/0904.1734" rel="nofollow noreferrer">this one</a> mostly on binary forms. In a nutshell, the above contraction can be encoded in a bipartite graph with 4 $\epsilon$-vertices and 4 $F$-vertices, all of degree three.</p> <p>Of course the above formalism generalizes with no difficulty (except managing notations) to plethysms of symmetric powers $S^p(S^q(\mathbb{C}^n))$ and similar $\Phi$, $\Psi$ maps giving the isomorphism with the relevant space of multisymmetric functions, i.e., $M_{[q,q,\ldots,q]}^{n}$ (with $q$ repeated $p$ times) or in simpler notation $M_{[q^p]}^{n}$. There is a canonical $GL_n$-equivariant map $$ H_{p,q}:S^p(S^q(\mathbb{C}^n))\rightarrow S^q(S^p(\mathbb{C}^n)) $$ with "H" standing for Hermite, Hadamard or Howe. I will not define it here but it's the first thing that would come to mind if asked to produce such a map (see <a href="https://arxiv.org/abs/1305.7387" rel="nofollow noreferrer">this review</a> by Landsberg for a detailed discussion in the wider context of geometric complexity theory). It is not hard to see that a homogeneous multisymmetric function in $M_{[p^q]}^{n}\simeq S^q(S^p(\mathbb{C}^n))$ is in the image of $H_{p,q}$ iff it is a polynomial in the homogeneous elementary multisymmetric functions which belong to $M_{[1^q]}^{n}$. In the context of the Aronhold invariant above, these elementary functions are the coefficients of the polynomial $$ P(x_1,x_2,x_3)=a_x b_x c_x d_x $$ and similarly in general.</p> <p>Finally, I can start discussing Will's excellent question. Let $T_{i_1,i_2,i_3}$ denote the tensor of interest with the action of $(g,h,k)\in (SL_n)^3$ given by $$ [(g,h,k)\cdot T]_{i_1,i_2,i_3}=\sum_{j_1,j_2,j_3=1}^{n} g_{i_1 j_1} h_{i_2 j_2} k_{i_3 j_3} T_{j_1,j_2,j_3}\ . $$ If Will's tensor had an even number $r$ of indices with the action of $(SL_n)^r$ the answer would be trivial. Take $n$ copies of $T$, contract all first indices of these copies with one $\epsilon$, then all second indices with another $\epsilon$, etc. For $r=3$ this tentative invariant vanishes identically. One can however keep the first indices free, and this gives a map $T\mapsto \Gamma(T)=F$ producing an $n$-ary $n$-ic $F(x_1,\ldots,x_n)$ corresponding to the symmetric tensor $$ F_{i_1\ldots i_n}=\sum_{j_1,\ldots,j_n, k_1\ldots,k_n=1}^{n} \epsilon_{j_1\ldots j_n} \epsilon_{k_1\ldots k_n} T_{i_1,j_1,k_1}\cdots T_{i_n,j_n,k_n}\ . $$ For symmetric $T$'s, such as Will's diagonal tensor $D_{ijk}=\delta_{ij}\delta_{ik}$, this is the Hessian covariant. Jason's idea can thus be rephrased as finding an $SL_n$-invariant $I=I(F)$ of $n$-ary $n$-ics $F$ and proposing $I(\Gamma(T))$ as an answer to Will's question. The main issue is to show $I(\Gamma(D))\neq 0$. Now $\Gamma(D)$ is the form $F(x_1,\ldots,x_n)=n!\ x_1 x_2\cdots x_n$.</p> <p>For $n$ even, the smallest degree for an invariant of $n$-ary $n$-ics $F(x_1,\ldots,x_n)$ is $n$ (unique up to scale). Its symbolic form is $$ \mathcal{S_n}=(a^{(1)} a^{(2)}\cdots a^{(n)})^n $$ where $a^{(1)}=(a_{1}^{(1)},a_{2}^{(1)}\ldots,a_{n}^{(1)}),\ldots,a^{(n)}=(a_{1}^{(n)},a_{2}^{(n)}\ldots,a_{n}^{(n)})$ are $n$ series of $n$ variables. Let $I_n=\Psi(\mathcal{S_n})$ denote this invariant. One can also write $$ I_n(F)=\epsilon_{i_{1,1}\ldots i_{1,n}}\epsilon_{i_{2,1}\ldots i_{2,n}}\cdots\epsilon_{i_{n,1}\ldots i_{n,n}} \ F_{i_{1,1}\ldots i_{n,1}}F_{i_{1,2}\ldots i_{n,2}}\cdots F_{i_{1,n}\ldots i_{n,n}} $$ with summation over indices understood. The evaluation of $I_n(\Gamma(D))$ is a multiple of the number of row-even minus the number of row-odd $n\times n$ Latin squares. Via work of Huang, Rota and Janssen showing that the result is nonzero is equivalent to the (widely open) Alon-Tarsi conjecture.</p> <p><strong>The following may be new (if not please let me know):</strong> For $n$ odd, the smallest degree for an invariant of $F(x_1,\ldots,x_n)$ is $n+1$ (unique up to scale). Its symbolic form is $$ \mathcal{S}=(a^{(2)} a^{(3)}\cdots a^{(n+1)})(a^{(1)} a^{(3)}\cdots a^{(n+1)})\cdots(a^{(1)} a^{(2)}\cdots a^{(n)})\ . $$ Again let me define $I_n=\Psi(\mathcal{S_n})$. This is nonzero because the polynomial in the $a^{(i)}$ is not only multihomogeneous of multidegree $[n^{n+1}]$ but also multisymmetric. For instance the transposition $a^{(1)}\leftrightarrow a^{(2)}$ exchanges the first two bracket factors while the other ones pick up a factor $(-1)$. This gives $(-1)^{n-1}=1$ since $n$ is odd. Other elementary transpositions $a^{(i)}\leftrightarrow a^{(i+1)}$ can be treated in the same way. Of course $I_3$ is Aronhold's degree 4 invariant. I don't know if $I_5$ could be of use for the study of quintic threefolds. I boldly (or foolishly?) make the following conjecture which I think is a natural analog of the Alon-Tarsi conjecture in the odd case.</p> <p><strong>Conjecture:</strong> For $n$ odd, $I_n$ does not vanish on $F(x_1,\ldots,x_n)=x_1 x_2\cdots x_n$.</p> <p>Here is a proof for $n=3$. The Aronhold invariant of degree 4 is the defining equation for the Veronese secant variety $\sigma_{3}(v_3(\mathbb{P}^2))$. Hence, it does not vanish on forms with border rank $&gt;3$. The border rank of the form $x_1 x_2\cdots x_n$ is exactly $2^{n-1}$ as shown by Oeding in <a href="https://arxiv.org/abs/1608.02530" rel="nofollow noreferrer">this article</a>. The case $n=3$ relevant to the present proof was shown earlier by <a href="https://arxiv.org/abs/0901.0487" rel="nofollow noreferrer">Landsberg and Teitler</a>. QED.</p> <p>A final remark regarding the endgame in Jason's argument with a view to producing an explicit invariant of moderate degree: it follows from what I mentioned above that saying that any homogeneous multisymmetric function in $M_{[p^q]}^{n}\simeq S^q(S^p(\mathbb{C}^n))$ can be expressed as a polynomial in the elementary ones is equivalent to the Hermite-Hadamard-Howe map $H_{p,q}$ being surjective and this is known for $p$ large with an explicit bound due to Brion (see the review by Landsberg I mentioned). The problem I was discussing is closer to the $p=q$ case related to the Foulkes-Howe conjecture. The latter is known to be false, but not by much! (see <a href="https://arxiv.org/abs/1509.03944" rel="nofollow noreferrer">this article</a>).</p> <p>Edit: I learned from C. Ikenmeyer that the conjecture about the nonvanishing of $I_n$ for $n$ odd is not new and is stated in Remark 3.26 in <a href="https://arxiv.org/abs/1511.02927" rel="nofollow noreferrer">https://arxiv.org/abs/1511.02927</a> However, I noticed recently that this is equivalent to the Alon-Tarsi conjecture for $n$ even. Indeed, the above invariants $I_n$ satisfy the following identity: $$ I_n(F(x_1,\ldots,x_{n-1})x_n)=(-1)^{\frac{n}{2}}\frac{n!}{n^n}\ I_{n-1}(F) $$ for all $n$ even and all forms $F$ which only depend on the first $n-1$ variables.</p> Answer 4: <p><B>Edit.</B> There was a mistake in the original formulation below of Theorem 2.4, p. 134 of Gelfand-Kapranov-Zelevinsky (now corrected). That theorem only applies after passing to the standard open affines. The upshot is that we know that there exists some integer $d_0&gt;0$ such that the determinant $D_{A,k}^{2d}$ gives a $\textbf{SL}(A)\times \textbf{SL}(B)\times \textbf{SL}(C)$-invariant polynomial that is nonzero on Will Sawin's tensor for all $d\geq d_0$. However, it is not immediately clear what exponent $d_0$ is necessary. An upper bound does follow from the Castelnuovo-Mumford regularity of the symmetric power $(\mathbb{P}(A))^n//\mathfrak{S}_n$ as a subvariety of $\mathbb{P}\text{Sym}^n_k(A)$, and there are bounds on this by Gotzmann. However, the Gotzmann bound is certainly too large for any effective computations.</p> <p><B>Original post.</B><br> I am just writing up my comments above as an answer. Let $k$ be a commutative, unital ring. Let $A$ be a free $k$-module of rank $n$. Denote by $\mathbb{P}(A)$ the Proj of the graded $k$-algebra $$S^\bullet_k(A^\vee) = \bigoplus_{d\geq 0} \text{Sym}^d_k(A^\vee).$$<br> For every integer $m\geq 0$, denote by $(S^\bullet_k(A^\vee))^{\otimes m}$ the $m$-fold tensor product $k$-algebra, $$(S^\bullet_k(A^\vee))^{\otimes m} = S^\bullet_k(A^\vee)\otimes_k \cdots \otimes_k S^\bullet_k(A^\vee).$$ This is naturally a $\mathbb{Z}^m$-graded algebra, and it has a natural action of the symmetric group $\mathfrak{S}_m$. Denote by $(S^\bullet_k(A^\vee))^{\otimes m}_{\langle 1,\dots,1 \rangle}$ the $\mathbb{Z}$-graded, $\mathfrak{S}_m$-fixed subalgebra, $$S^\bullet_k(A^\vee))^{\otimes m}_{\langle 1,\dots, 1 \rangle} = \bigoplus_{d\geq 0} \text{Sym}^d_k(A^\vee)\otimes_k \cdots \otimes_k \text{Sym}^d_k(A^\vee).$$ Now assume that $m!$ is invertible in $k$, so that $\text{Sym}^m_k(A)^\vee$ is canonically isomorphic to $\text{Sym}^m_k(A^\vee)$ as $k$-vector spaces. Then the transpose of the $m$-multilinear $k$-module homomorphism, $$A^m \to \text{Sym}^m_k(A), \ \ (a_1,\dots,a_m)\mapsto a_1\cdots a_m,$$ induces a $k$-algebra homomorphism, $$S_k^\bullet(\text{Sym}^m_k(A^\vee)) \to (S^\bullet_k(A^\vee))^{\otimes m}_{\langle 1,\dots,1 \rangle}.$$ This, in turn, defines a $k$-morphism, $$\gamma': (\mathbb{P}(A))^m\to \mathbb{P}(\text{Sym}^m_k(A)),\ \ ([a_1],\dots,[a_m]) \mapsto [a_1\cdots a_m]$$ where $(\mathbb{P}(A))^m$ is shorthand for the $m$-fold self-fiber product $\mathbb{P}(A)\times_{\text{Spec}\ k}\cdots \times_{\text{Spec}\ k}\mathbb{P}(A)$. The morphism $\gamma'$ is invariant for the natural action of the symmetric group $\mathfrak{S}_m$ on $(\mathbb{P}(A))^m$. Thus, $\gamma'$ factors through the quotient of this group action. Denote by $\gamma$ the induced $k$-morphism, $$\gamma: (\mathbb{P}(A))^m//\mathfrak{S}_m \to \mathbb{P}(\text{Sym}^m_k(A)).$$</p> <p>It is a classical fact that $\gamma$ is a closed immersion. This is explicitly stated and proved in Proposition 2.1 and Theorem 2.2, Chapter 4, pp. 132-133 of Gelfand-Kapranov-Zelevinsky. In fact, Gelfand-Kapranov-Zelevinsky prove much more: for every nonzero $x_i\in A^\vee$, denoting by $\gamma_m(x_i)\in \text{Sym}^m_k(A^\vee)$ the associated $m^{\text{th}}$ power, then the induced $k$-algebra homomorphism of affine opens, $$S_k^\bullet(\text{Sym}^m_k(A^\vee))[1/\gamma_m(x_i)] \to [(S_k^\bullet(A^\vee))^{\otimes m}_{\langle 1,\dots,1 \rangle}]^{\mathfrak{S}_m}[1/(x_{1,i}\cdots x_{m,i})],$$ is a surjective $k$-algebra homomorphism. This is essentially Theorem 2.4 on p. 134. From this it follows that for every $\mathfrak{S}_m$-invariant, homogeneous element $D$ of degree $r&gt;0$ in $(S_k^\bullet(A^\vee))^{\otimes m}_{(r,\dots,r)}$, there exists an integer $e_0$ such that for every $e\geq e_0$, $D$ times the image of $\text{Sym}^e_k(\text{Sym}^m_k(A^\vee))$ is in the image of $\text{Sym}^{e+r}_k(\text{Sym}^m_k(A^\vee))$. Thus there exists an integer $d&gt;0$ such that $D^d$ is in the image of $\text{Sym}^{dr}_k(\text{Sym}^m_k(A^\vee))$.</p> <p>In particular, for $m=n$ (assuming that $n!$ is invertible in $k$), the determinant map, $$\text{det}_{A,k}: A^n \to \bigwedge^n_k(A),$$ defines an element $D_{A,k} \in (S_k^\bullet(A^\vee))^{\otimes n}_{(1,\dots,1)}$. This element is <B>not</B> $\mathfrak{S}_m$-invariant. However, the square, $D_{A,k}^2$ is $\mathfrak{S}_m$-invariant. Thus, $D_{A,k}^{2d}$ is in the image $I_{2d}$ of $\text{Sym}_k^{2d}(\text{Sym}_k^n(A^\vee))$ for some integer $d&gt;0$.</p> <p>Finally, if we assume that $k$ is a field of characteristic $0$, then $\textbf{SL}(A)$ is linearly reductive. Thus, the induced $k$-linear map, $$\text{Sym}^{2d}_k(\text{Sym}^m_k(A^\vee))^{\textbf{SL}(A)} \to I_{2d}^{\textbf{SL}(A)},$$ is surjective. In particular, the $\textbf{SL}(A)$-invariant element $D_{A,k}^{2d}$ is in the image of $\left( \text{Sym}_k^{2d}(\text{Sym}_k^n(A^\vee)) \right)^{\textbf{SL}(A)}$. So there exists some $\textbf{SL}(A)$-invariant element $\Delta_{A,k}$ mappsing to $D_{A,k}^{2d}$.</p> <p>Now put this together with the tensor operation from the comments. For $k$-vector spaces $A$, $B$, and $C$ of dimension $n$, for every $k$-linear map, $$t:A^\vee \to \text{Hom}_k(B,C),$$ there is an associated element $$\widetilde{t} \in \text{Sym}^n(A) \otimes_k \text{Hom}_k(\bigwedge^n_k(B),\bigwedge^n_k(C)),$$ giving the polynomial map, $$ A^\vee\to \text{Hom}_k(\bigwedge^n_k(B),\bigwedge^n_k(C)), \ \ a \mapsto \text{det}(t(a)).$$ The associated map $$\text{Hom}_k(A^\vee,\text{Hom}_k(B,C))\to \text{Sym}^n(A)\otimes_k \text{Hom}_k(\bigwedge^n_k(B),\bigwedge^n_k(C)), \ \ t\mapsto \widetilde{t},$$ is a degree $n$, homogeneous $k$-polynomial map that is equivariant for the natural actions of $\textbf{SL}(A)\times \textbf{SL}(B)\times \textbf{SL}(C)$. Via the duality of $\text{Sym}^n(A)$ and $\text{Sym}^n(A^\vee)$, the polynomial $\Delta_{A,k}$ determines a $\textbf{SL}(A)\times \textbf{SL}(B)\times \textbf{SL}(C)$-invariant homogeneous polynomial of degree $2dn$ on $\text{Hom}_k(A^\vee,\text{Hom}_k(B,C))$. </p> <p>For the specific tensor $t$ that you wrote down, with respect to appropriate bases for $A$, $B$, and $C$, for every $(a_1,\dots,a_n)\in A$, $t(a_1,\dots,a_n)$ is the linear transformation whose matrix representative is the diagonal $n\times n$ matrix with entries $(a_1,\dots,a_n)$. Thus, $\widetilde{t}$ equals $(a_1\cdots a_n)\otimes (b_1\wedge \dots \wedge b_n)\otimes (c_1\wedge \dots c_n)$. In particular, since $a_1\cdots a_n$ is in the image of $\gamma$, $\Delta_{A,k}$ restricts on $\widetilde{t}$ as a power of the determinant. Since $a_1\cdots a_n$ is a product of $n$ linearly independent linear polynomials, the determinant is nonzero. Thus, $\Delta_{A,k}$ is a $\textbf{SL}(A)\times \textbf{SL}(B)\times \textbf{SL}(C)$-invariant, homogeneous polynomial of degree $2n$ that is nonzero on your tensor. </p> Answer 5: <p>Let me start with some remarks about the classical symbolic method (without which one cannot understand 19th century invariant theory) and multisymmetric functions. I will use an example first. Take four series of three variables $a=(a_1,a_2,a_3)$, $b=(b_1,b_2,b_3)$, $c=(c_1,c_2,c_3)$ and $d=(d_1,d_2,d_3)$. Now define the polynomial $\mathcal{A}$ in these 12 indeterminates given by $$ \mathcal{A}=(bcd)(acd)(abd)(abc) $$ where we used a shorthand notation for determinants $$ (bcd)=\left| \begin{array}{ccc} b_1 &amp; b_2 &amp; b_3 \\ c_1 &amp; c_2 &amp; c_3 \\ d_1 &amp; d_2 &amp; d_3 \\ \end{array} \right| $$ and similarly for the other "bracket factors" $(acd)$, etc. This is an example of multisymmetric function in the "letters" $a,b,c,d$ which is multihomogeneous of multidegree $[3,3,3,3]$. Let me write $M_{[3,3,3,3]}^3$ for the space of such multisymmetric functions. The subscript is the multidegree while the superscript refers to the number of variables in each series. The symmetric group acting is $\mathfrak{S}_4$ which permutes these series. If each series had one variable in it instead of three we would be in the familiar realm of the theory of ordinary symmetric functions. The analogous theory of multisymmetric functions is much more complicated and has been studied classically by Euler, Lagrange, Poisson, Schl&auml;fli, Brill, Gordan, Junker and more recently by Angeniol, Dalbec, Briand, Vaccarino, Domokos and others. A nice yet old reference on this is the book on elimination theory by Faa di Bruno I mentioned in <a href="https://mathoverflow.net/questions/117415/old-books-still-used/117417#117417">this MO answer</a>.</p> <p>Consider now a generic ternary cubic form $F=F(x_1,x_2,x_3)$. One can write it as $$ F(x)=\sum_{i_1,i_2,i_3=1}^3 F_{i_1,i_2,i_3} x_{i_1} x_{i_2} x_{i_3} $$ where $(F_{i_1,i_2,i_3})$ is a symmetric tensor (by which I mean a "matrix" rather than a Bourbaki tensor). This form lives in the symmetric power $S^3(\mathbb{C}^3)$ but I will be (intentionally) sloppy and write $S^3$ instead. The space of degree 4 homogeneous polynomials in the coefficients of $F$ will be denoted by $S^4(S^3)$.</p> <p>Let me define a linear map $\Phi:S^4(S^3)\rightarrow M_{[3,3,3,3]}^3$ as follows. Given a degree four homogeneous polynomial $C(F)$ first construct its full polarization $$ \widetilde{C}(F_1,F_2,F_3,F_4)=\frac{1}{4!} \frac{\partial^4}{\partial t_1\partial t_2\partial t_3\partial t_4} C(t_1F_1+t_2 F_2+t_3 F_3+t_4 F_4) $$ which satisfies $\widetilde{C}(F,F,F,F)=C(F)$. Now set $F_1(x)=a_{x}^3$ where $a_x$ is the classical notation for the linear form $a_1x_1+a_2x_2+a_3x_3$ and similarly specialize to the cubes of linear forms $F_2(x)=b_{x}^3$, $F_3(x)=c_{x}^3$ and $F_4(x)=d_{x}^3$. Finally, define $\Phi(C)=\widetilde{C}(F_1,F_2,F_3,F_4)$.</p> <p>I will also define another linear map $\Psi: M_{[3,3,3,3]}^3\rightarrow S^4(S^3)$. Let $\mathcal{S}=\mathcal{S}(a,b,c,d)$ be a multisymmetric function in $M_{[3,3,3,3]}^3$. We define the differential operator $$ \mathcal{D}=\frac{1}{3!^4} F\left(\frac{\partial}{\partial a_1},\frac{\partial}{\partial a_2},\frac{\partial}{\partial a_3}\right) F\left(\frac{\partial}{\partial b_1},\frac{\partial}{\partial b_2},\frac{\partial}{\partial b_3}\right) $$ $$ F\left(\frac{\partial}{\partial c_1},\frac{\partial}{\partial c_2},\frac{\partial}{\partial c_3}\right) F\left(\frac{\partial}{\partial d_1},\frac{\partial}{\partial d_2},\frac{\partial}{\partial d_3}\right) $$ and let $\Psi(\mathcal{S})=\mathcal{D}\ \mathcal{S}$, i.e., the result of applying the differential operator $\mathcal{D}$ to the polynomial $\mathcal{S}$.</p> <p>It is not hard to see that $\Phi$ and $\Psi$ are inverses of one another. In the particular case of $\mathcal{A}$ above, the image $\Psi(\mathcal{A})$ is Aronhold's degree four $SL_3$ invariant of plane cubics and $\mathcal{A}$ itself is called the (German) symbolic form of that invariant. The classics usually omitted "$\Psi(\cdots)$" but they knew perfectly well what they were doing. Also note that this formalism, in an essentially equivalent form, was invented earlier by Cayley as I briefly explained in Section 1.11.4 of my article <a href="https://arxiv.org/abs/1604.05259" rel="nofollow noreferrer">"A Second-Quantized Kolmogorov-Chentsov Theorem"</a>. One can easily write $\Psi(\mathcal{A})$ applied to $F$, in "physicsy" tensor contraction notation as $$ \epsilon_{i_{1} i_{2} i_{3}}\epsilon_{i_{4} i_{5} i_{6}}\epsilon_{i_{7} i_{8} i_{9}}\epsilon_{i_{10} i_{11} i_{12}} F_{i_{4} i_{7} i_{10}} F_{i_{1} i_{8} i_{11}} F_{i_{2} i_{5} i_{12}} F_{i_{3} i_{6} i_{9}} $$ with summation over all twelve indices in the set $\{1,2,3\}$. I used the standard physics notation for the Levi-Civita tensor $\epsilon_{i_1 i_2 i_3}$ equal to zero if some indices are equal and otherwise to the sign of the permutation $i_1,i_2,i_3$ of $1,2,3$. This notation is consistent with the one I used in my previous answers to these MO questions: <a href="https://mathoverflow.net/questions/121715/invariant-polynomials-for-a-product-of-algebraic-groups">Q1</a>, <a href="https://mathoverflow.net/questions/257709/ring-of-invariants-of-operatornamesl-6-acting-on-lambda3-mathbb-c6">Q2</a>, and <a href="https://mathoverflow.net/questions/255492/how-to-constructively-combinatorially-prove-schur-weyl-duality">Q3</a>. At bottom, this is an intrinsically diagrammatic language and if the reader had trouble following what I said so far, it's probably because I did not draw any pictures or "finite-dimensional Feynman diagrams" (I don't know how to do this on MO). To see these pictures, please have a look at <a href="https://arxiv.org/abs/1402.2669" rel="nofollow noreferrer">this article</a> about an invariant of cubic threefolds or Section 2 of <a href="https://arxiv.org/abs/0904.1734" rel="nofollow noreferrer">this one</a> mostly on binary forms. In a nutshell, the above contraction can be encoded in a bipartite graph with 4 $\epsilon$-vertices and 4 $F$-vertices, all of degree three.</p> <p>Of course the above formalism generalizes with no difficulty (except managing notations) to plethysms of symmetric powers $S^p(S^q(\mathbb{C}^n))$ and similar $\Phi$, $\Psi$ maps giving the isomorphism with the relevant space of multisymmetric functions, i.e., $M_{[q,q,\ldots,q]}^{n}$ (with $q$ repeated $p$ times) or in simpler notation $M_{[q^p]}^{n}$. There is a canonical $GL_n$-equivariant map $$ H_{p,q}:S^p(S^q(\mathbb{C}^n))\rightarrow S^q(S^p(\mathbb{C}^n)) $$ with "H" standing for Hermite, Hadamard or Howe. I will not define it here but it's the first thing that would come to mind if asked to produce such a map (see <a href="https://arxiv.org/abs/1305.7387" rel="nofollow noreferrer">this review</a> by Landsberg for a detailed discussion in the wider context of geometric complexity theory). It is not hard to see that a homogeneous multisymmetric function in $M_{[p^q]}^{n}\simeq S^q(S^p(\mathbb{C}^n))$ is in the image of $H_{p,q}$ iff it is a polynomial in the homogeneous elementary multisymmetric functions which belong to $M_{[1^q]}^{n}$. In the context of the Aronhold invariant above, these elementary functions are the coefficients of the polynomial $$ P(x_1,x_2,x_3)=a_x b_x c_x d_x $$ and similarly in general.</p> <p>Finally, I can start discussing Will's excellent question. Let $T_{i_1,i_2,i_3}$ denote the tensor of interest with the action of $(g,h,k)\in (SL_n)^3$ given by $$ [(g,h,k)\cdot T]_{i_1,i_2,i_3}=\sum_{j_1,j_2,j_3=1}^{n} g_{i_1 j_1} h_{i_2 j_2} k_{i_3 j_3} T_{j_1,j_2,j_3}\ . $$ If Will's tensor had an even number $r$ of indices with the action of $(SL_n)^r$ the answer would be trivial. Take $n$ copies of $T$, contract all first indices of these copies with one $\epsilon$, then all second indices with another $\epsilon$, etc. For $r=3$ this tentative invariant vanishes identically. One can however keep the first indices free, and this gives a map $T\mapsto \Gamma(T)=F$ producing an $n$-ary $n$-ic $F(x_1,\ldots,x_n)$ corresponding to the symmetric tensor $$ F_{i_1\ldots i_n}=\sum_{j_1,\ldots,j_n, k_1\ldots,k_n=1}^{n} \epsilon_{j_1\ldots j_n} \epsilon_{k_1\ldots k_n} T_{i_1,j_1,k_1}\cdots T_{i_n,j_n,k_n}\ . $$ For symmetric $T$'s, such as Will's diagonal tensor $D_{ijk}=\delta_{ij}\delta_{ik}$, this is the Hessian covariant. Jason's idea can thus be rephrased as finding an $SL_n$-invariant $I=I(F)$ of $n$-ary $n$-ics $F$ and proposing $I(\Gamma(T))$ as an answer to Will's question. The main issue is to show $I(\Gamma(D))\neq 0$. Now $\Gamma(D)$ is the form $F(x_1,\ldots,x_n)=n!\ x_1 x_2\cdots x_n$.</p> <p>For $n$ even, the smallest degree for an invariant of $n$-ary $n$-ics $F(x_1,\ldots,x_n)$ is $n$ (unique up to scale). Its symbolic form is $$ \mathcal{S_n}=(a^{(1)} a^{(2)}\cdots a^{(n)})^n $$ where $a^{(1)}=(a_{1}^{(1)},a_{2}^{(1)}\ldots,a_{n}^{(1)}),\ldots,a^{(n)}=(a_{1}^{(n)},a_{2}^{(n)}\ldots,a_{n}^{(n)})$ are $n$ series of $n$ variables. Let $I_n=\Psi(\mathcal{S_n})$ denote this invariant. One can also write $$ I_n(F)=\epsilon_{i_{1,1}\ldots i_{1,n}}\epsilon_{i_{2,1}\ldots i_{2,n}}\cdots\epsilon_{i_{n,1}\ldots i_{n,n}} \ F_{i_{1,1}\ldots i_{n,1}}F_{i_{1,2}\ldots i_{n,2}}\cdots F_{i_{1,n}\ldots i_{n,n}} $$ with summation over indices understood. The evaluation of $I_n(\Gamma(D))$ is a multiple of the number of row-even minus the number of row-odd $n\times n$ Latin squares. Via work of Huang, Rota and Janssen showing that the result is nonzero is equivalent to the (widely open) Alon-Tarsi conjecture.</p> <p><strong>The following may be new (if not please let me know):</strong> For $n$ odd, the smallest degree for an invariant of $F(x_1,\ldots,x_n)$ is $n+1$ (unique up to scale). Its symbolic form is $$ \mathcal{S}=(a^{(2)} a^{(3)}\cdots a^{(n+1)})(a^{(1)} a^{(3)}\cdots a^{(n+1)})\cdots(a^{(1)} a^{(2)}\cdots a^{(n)})\ . $$ Again let me define $I_n=\Psi(\mathcal{S_n})$. This is nonzero because the polynomial in the $a^{(i)}$ is not only multihomogeneous of multidegree $[n^{n+1}]$ but also multisymmetric. For instance the transposition $a^{(1)}\leftrightarrow a^{(2)}$ exchanges the first two bracket factors while the other ones pick up a factor $(-1)$. This gives $(-1)^{n-1}=1$ since $n$ is odd. Other elementary transpositions $a^{(i)}\leftrightarrow a^{(i+1)}$ can be treated in the same way. Of course $I_3$ is Aronhold's degree 4 invariant. I don't know if $I_5$ could be of use for the study of quintic threefolds. I boldly (or foolishly?) make the following conjecture which I think is a natural analog of the Alon-Tarsi conjecture in the odd case.</p> <p><strong>Conjecture:</strong> For $n$ odd, $I_n$ does not vanish on $F(x_1,\ldots,x_n)=x_1 x_2\cdots x_n$.</p> <p>Here is a proof for $n=3$. The Aronhold invariant of degree 4 is the defining equation for the Veronese secant variety $\sigma_{3}(v_3(\mathbb{P}^2))$. Hence, it does not vanish on forms with border rank $&gt;3$. The border rank of the form $x_1 x_2\cdots x_n$ is exactly $2^{n-1}$ as shown by Oeding in <a href="https://arxiv.org/abs/1608.02530" rel="nofollow noreferrer">this article</a>. The case $n=3$ relevant to the present proof was shown earlier by <a href="https://arxiv.org/abs/0901.0487" rel="nofollow noreferrer">Landsberg and Teitler</a>. QED.</p> <p>A final remark regarding the endgame in Jason's argument with a view to producing an explicit invariant of moderate degree: it follows from what I mentioned above that saying that any homogeneous multisymmetric function in $M_{[p^q]}^{n}\simeq S^q(S^p(\mathbb{C}^n))$ can be expressed as a polynomial in the elementary ones is equivalent to the Hermite-Hadamard-Howe map $H_{p,q}$ being surjective and this is known for $p$ large with an explicit bound due to Brion (see the review by Landsberg I mentioned). The problem I was discussing is closer to the $p=q$ case related to the Foulkes-Howe conjecture. The latter is known to be false, but not by much! (see <a href="https://arxiv.org/abs/1509.03944" rel="nofollow noreferrer">this article</a>).</p> <p>Edit: I learned from C. Ikenmeyer that the conjecture about the nonvanishing of $I_n$ for $n$ odd is not new and is stated in Remark 3.26 in <a href="https://arxiv.org/abs/1511.02927" rel="nofollow noreferrer">https://arxiv.org/abs/1511.02927</a> However, I noticed recently that this is equivalent to the Alon-Tarsi conjecture for $n$ even. Indeed, the above invariants $I_n$ satisfy the following identity: $$ I_n(F(x_1,\ldots,x_{n-1})x_n)=(-1)^{\frac{n}{2}}\frac{n!}{n^n}\ I_{n-1}(F) $$ for all $n$ even and all forms $F$ which only depend on the first $n-1$ variables.</p> Answer 6: <p><B>Edit.</B> There was a mistake in the original formulation below of Theorem 2.4, p. 134 of Gelfand-Kapranov-Zelevinsky (now corrected). That theorem only applies after passing to the standard open affines. The upshot is that we know that there exists some integer $d_0&gt;0$ such that the determinant $D_{A,k}^{2d}$ gives a $\textbf{SL}(A)\times \textbf{SL}(B)\times \textbf{SL}(C)$-invariant polynomial that is nonzero on Will Sawin's tensor for all $d\geq d_0$. However, it is not immediately clear what exponent $d_0$ is necessary. An upper bound does follow from the Castelnuovo-Mumford regularity of the symmetric power $(\mathbb{P}(A))^n//\mathfrak{S}_n$ as a subvariety of $\mathbb{P}\text{Sym}^n_k(A)$, and there are bounds on this by Gotzmann. However, the Gotzmann bound is certainly too large for any effective computations.</p> <p><B>Original post.</B><br> I am just writing up my comments above as an answer. Let $k$ be a commutative, unital ring. Let $A$ be a free $k$-module of rank $n$. Denote by $\mathbb{P}(A)$ the Proj of the graded $k$-algebra $$S^\bullet_k(A^\vee) = \bigoplus_{d\geq 0} \text{Sym}^d_k(A^\vee).$$<br> For every integer $m\geq 0$, denote by $(S^\bullet_k(A^\vee))^{\otimes m}$ the $m$-fold tensor product $k$-algebra, $$(S^\bullet_k(A^\vee))^{\otimes m} = S^\bullet_k(A^\vee)\otimes_k \cdots \otimes_k S^\bullet_k(A^\vee).$$ This is naturally a $\mathbb{Z}^m$-graded algebra, and it has a natural action of the symmetric group $\mathfrak{S}_m$. Denote by $(S^\bullet_k(A^\vee))^{\otimes m}_{\langle 1,\dots,1 \rangle}$ the $\mathbb{Z}$-graded, $\mathfrak{S}_m$-fixed subalgebra, $$S^\bullet_k(A^\vee))^{\otimes m}_{\langle 1,\dots, 1 \rangle} = \bigoplus_{d\geq 0} \text{Sym}^d_k(A^\vee)\otimes_k \cdots \otimes_k \text{Sym}^d_k(A^\vee).$$ Now assume that $m!$ is invertible in $k$, so that $\text{Sym}^m_k(A)^\vee$ is canonically isomorphic to $\text{Sym}^m_k(A^\vee)$ as $k$-vector spaces. Then the transpose of the $m$-multilinear $k$-module homomorphism, $$A^m \to \text{Sym}^m_k(A), \ \ (a_1,\dots,a_m)\mapsto a_1\cdots a_m,$$ induces a $k$-algebra homomorphism, $$S_k^\bullet(\text{Sym}^m_k(A^\vee)) \to (S^\bullet_k(A^\vee))^{\otimes m}_{\langle 1,\dots,1 \rangle}.$$ This, in turn, defines a $k$-morphism, $$\gamma': (\mathbb{P}(A))^m\to \mathbb{P}(\text{Sym}^m_k(A)),\ \ ([a_1],\dots,[a_m]) \mapsto [a_1\cdots a_m]$$ where $(\mathbb{P}(A))^m$ is shorthand for the $m$-fold self-fiber product $\mathbb{P}(A)\times_{\text{Spec}\ k}\cdots \times_{\text{Spec}\ k}\mathbb{P}(A)$. The morphism $\gamma'$ is invariant for the natural action of the symmetric group $\mathfrak{S}_m$ on $(\mathbb{P}(A))^m$. Thus, $\gamma'$ factors through the quotient of this group action. Denote by $\gamma$ the induced $k$-morphism, $$\gamma: (\mathbb{P}(A))^m//\mathfrak{S}_m \to \mathbb{P}(\text{Sym}^m_k(A)).$$</p> <p>It is a classical fact that $\gamma$ is a closed immersion. This is explicitly stated and proved in Proposition 2.1 and Theorem 2.2, Chapter 4, pp. 132-133 of Gelfand-Kapranov-Zelevinsky. In fact, Gelfand-Kapranov-Zelevinsky prove much more: for every nonzero $x_i\in A^\vee$, denoting by $\gamma_m(x_i)\in \text{Sym}^m_k(A^\vee)$ the associated $m^{\text{th}}$ power, then the induced $k$-algebra homomorphism of affine opens, $$S_k^\bullet(\text{Sym}^m_k(A^\vee))[1/\gamma_m(x_i)] \to [(S_k^\bullet(A^\vee))^{\otimes m}_{\langle 1,\dots,1 \rangle}]^{\mathfrak{S}_m}[1/(x_{1,i}\cdots x_{m,i})],$$ is a surjective $k$-algebra homomorphism. This is essentially Theorem 2.4 on p. 134. From this it follows that for every $\mathfrak{S}_m$-invariant, homogeneous element $D$ of degree $r&gt;0$ in $(S_k^\bullet(A^\vee))^{\otimes m}_{(r,\dots,r)}$, there exists an integer $e_0$ such that for every $e\geq e_0$, $D$ times the image of $\text{Sym}^e_k(\text{Sym}^m_k(A^\vee))$ is in the image of $\text{Sym}^{e+r}_k(\text{Sym}^m_k(A^\vee))$. Thus there exists an integer $d&gt;0$ such that $D^d$ is in the image of $\text{Sym}^{dr}_k(\text{Sym}^m_k(A^\vee))$.</p> <p>In particular, for $m=n$ (assuming that $n!$ is invertible in $k$), the determinant map, $$\text{det}_{A,k}: A^n \to \bigwedge^n_k(A),$$ defines an element $D_{A,k} \in (S_k^\bullet(A^\vee))^{\otimes n}_{(1,\dots,1)}$. This element is <B>not</B> $\mathfrak{S}_m$-invariant. However, the square, $D_{A,k}^2$ is $\mathfrak{S}_m$-invariant. Thus, $D_{A,k}^{2d}$ is in the image $I_{2d}$ of $\text{Sym}_k^{2d}(\text{Sym}_k^n(A^\vee))$ for some integer $d&gt;0$.</p> <p>Finally, if we assume that $k$ is a field of characteristic $0$, then $\textbf{SL}(A)$ is linearly reductive. Thus, the induced $k$-linear map, $$\text{Sym}^{2d}_k(\text{Sym}^m_k(A^\vee))^{\textbf{SL}(A)} \to I_{2d}^{\textbf{SL}(A)},$$ is surjective. In particular, the $\textbf{SL}(A)$-invariant element $D_{A,k}^{2d}$ is in the image of $\left( \text{Sym}_k^{2d}(\text{Sym}_k^n(A^\vee)) \right)^{\textbf{SL}(A)}$. So there exists some $\textbf{SL}(A)$-invariant element $\Delta_{A,k}$ mappsing to $D_{A,k}^{2d}$.</p> <p>Now put this together with the tensor operation from the comments. For $k$-vector spaces $A$, $B$, and $C$ of dimension $n$, for every $k$-linear map, $$t:A^\vee \to \text{Hom}_k(B,C),$$ there is an associated element $$\widetilde{t} \in \text{Sym}^n(A) \otimes_k \text{Hom}_k(\bigwedge^n_k(B),\bigwedge^n_k(C)),$$ giving the polynomial map, $$ A^\vee\to \text{Hom}_k(\bigwedge^n_k(B),\bigwedge^n_k(C)), \ \ a \mapsto \text{det}(t(a)).$$ The associated map $$\text{Hom}_k(A^\vee,\text{Hom}_k(B,C))\to \text{Sym}^n(A)\otimes_k \text{Hom}_k(\bigwedge^n_k(B),\bigwedge^n_k(C)), \ \ t\mapsto \widetilde{t},$$ is a degree $n$, homogeneous $k$-polynomial map that is equivariant for the natural actions of $\textbf{SL}(A)\times \textbf{SL}(B)\times \textbf{SL}(C)$. Via the duality of $\text{Sym}^n(A)$ and $\text{Sym}^n(A^\vee)$, the polynomial $\Delta_{A,k}$ determines a $\textbf{SL}(A)\times \textbf{SL}(B)\times \textbf{SL}(C)$-invariant homogeneous polynomial of degree $2dn$ on $\text{Hom}_k(A^\vee,\text{Hom}_k(B,C))$. </p> <p>For the specific tensor $t$ that you wrote down, with respect to appropriate bases for $A$, $B$, and $C$, for every $(a_1,\dots,a_n)\in A$, $t(a_1,\dots,a_n)$ is the linear transformation whose matrix representative is the diagonal $n\times n$ matrix with entries $(a_1,\dots,a_n)$. Thus, $\widetilde{t}$ equals $(a_1\cdots a_n)\otimes (b_1\wedge \dots \wedge b_n)\otimes (c_1\wedge \dots c_n)$. In particular, since $a_1\cdots a_n$ is in the image of $\gamma$, $\Delta_{A,k}$ restricts on $\widetilde{t}$ as a power of the determinant. Since $a_1\cdots a_n$ is a product of $n$ linearly independent linear polynomials, the determinant is nonzero. Thus, $\Delta_{A,k}$ is a $\textbf{SL}(A)\times \textbf{SL}(B)\times \textbf{SL}(C)$-invariant, homogeneous polynomial of degree $2n$ that is nonzero on your tensor. </p>
Putting algebraic curves in $\mathbb{R}^3$
259,526
18
https://mathoverflow.net/questions/259526/putting-algebraic-curves-in-mathbbr3
<p>Let $X \subset \mathbb{C}^2$ be a smooth algebraic curve. Thinking of $\mathbb{C}^2$ as $\mathbb{R}^4$, is there a smooth map $\phi: \mathbb{C}^2 \to \mathbb{R}^3$ so that $\phi: X \to \mathbb{R}^3$ is a closed injective immersion? I mean this question in two senses:</p> <ul> <li><p>Does the map $\phi$ exist at all?</p></li> <li><p>Can we actually write down some simple formula for it, in terms of coordinates $x_1+i y_1$, $x_2+i y_2$ on $\mathbb{C}^2$ and the defining equation $F(x_1+iy_1,x_2+iy_2)=0$ of $X$?</p></li> </ul> <p>The motivation for this question is that I'd like to make some pretty pictures of algebraic curves.</p>
[ { "answer_id": 259603, "score": 4, "text": "<p>Taking this as two questions, the second being more interesting than the first, I can at least answer the first (less interesting) question. The answer is 'Yes, such maps $\\phi:\\mathbb{C}^2\\to\\mathbb{R}^3$ do exist, though they (necessarily) depend on the smooth curve $X$'.</p>\n\n<p>Given a smooth algebraic curve $X\\subset\\mathbb{C}^2$, suppose that $X$ is defined as the zero locus of an reduced polynomial $F(z,w)$ (i.e., $F$ has no multiple factors). Then, by the assumption that $X$ is smooth (by which I assume that David means 'smooth, embedded'), the polynomials, $F$, $F_z$ and $F_w$ have no common zeros. In fact, for each $x\\in X$, the vector $N(x)=\\bigl(\\,\\overline{F_z(x)},\\,\\overline{F_w(x)}\\,\\bigr)$ is nonzero and (unitarily) orthogonal to the (complex) tangent line to $X$ at $x$, since the $1$-form $\\mathrm{d}F = F_z\\,\\mathrm{d}z+F_w\\,\\mathrm{d}w$ vanishes when pulled back to the curve $X$. Let $U(x)= N(x)/|N(x)|$ be the corresponding unit vector.</p>\n\n<p>Now, because $X$ is smooth and algebraic, outside a compact set, it is asymptotic to a finite set of lines, and it is not difficult to see that there is a positive function $e:X\\to (0,1)$ such that the mapping $S:X\\times\\Delta(1)\\to\\mathbb{C}^2$ (where $\\Delta(r)\\subset\\mathbb{C}$ is the disk of radius $r&gt;0$ about $0$) defined by\n$$\nS(x,t) = x + t\\,e(x)U(x)\n$$\nis an injective diffeomorphism. (If all of the asmptotic lines of $X$ are distinct, one can even take $e$ to be a (suitably small) constant.)</p>\n\n<p>Now, we also know that $X$ is a compact (oriented) Riemann surface with a (nonzero) finite number of points removed. (In fact, $X$ has no compact components.) As such, there exists a smooth, closed embedding $\\psi: X\\to \\mathbb{R}^3$ with the property that the normal 'tube' of radius $1$ around $\\psi(X)$ is also smoothly embedded. Let $u:X\\to S^2$ be a unit normal vector field for the immersion $\\psi$ and extend $\\psi$ to $\\psi:X\\times [-\\tfrac12,\\tfrac12]\\to\\mathbb{R}^3$ by setting\n$$\n\\psi(x,t) = \\psi(x) + t\\,u(x)\n$$\nfor $|t|\\le \\tfrac12$. Now, I claim that there is a smooth map $\\phi:S\\bigl(X\\times \\Delta(\\tfrac12)\\bigr)\\to \\mathbb{R}^3$ that satisfies\n$$\n\\phi\\bigl(S(x,t)\\bigr) = \\psi\\bigl(x,\\mathrm{Re}(t)\\bigr)\n$$\nwhenever $|t|\\le\\tfrac12$ and that it is a smooth submersion on an open neighborhood of $X\\subset S\\bigl(X\\times \\Delta(\\tfrac12)\\bigr)$ that is injective and immersive on $X$ itself.</p>\n\n<p>Now, extend $\\phi$ smoothly any way one likes beyond the set $S\\bigl(X\\times \\Delta(\\tfrac12)\\bigr)\\subset\\mathbb{C}^2$. (One can even require that $\\phi$ take the complement of $S\\bigl(X\\times \\Delta(1)\\bigr)$ to a single point of $\\mathbb{R}^3$.)</p>\n\n<p>Unfortunately, there may be no simple recipe for choosing $\\psi$, which is what one would really need to get an affirmative answer to the second question.</p>\n" }, { "answer_id": 259607, "score": 1, "text": "<p>Henry Segerman has asked a similar <a href=\"https://mathoverflow.net/questions/200172/map-of-the-klein-quartic-from-cp2-to-r3\">question</a> about curves in $\\mathbb{CP}^2$ and then managed to produce the <a href=\"http://gallery.bridgesmathart.org/exhibitions/2016-joint-mathematics-meetings/schleimer-segerman\" rel=\"nofollow noreferrer\">sculpture</a> (of the Klein quartic). So perhaps he will have something interesting to say...</p>\n" }, { "answer_id": 260495, "score": 1, "text": "<p>Consider the surface in $\\mathbb{R}^4$ with equations \n\\begin{align*}\n x_1^2+x_2^2+x_3^2+x_4^2 &amp;=1 \\\\\n(3x_3^2-2)x_4-\\sqrt{2}(x_1^2-x_2^2)x_3 &amp;=0.\n\\end{align*} \n<a href=\"https://i.stack.imgur.com/4E9WV.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/4E9WV.png\" alt=\"enter image description here\"></a></p>\n\n<p>This is conformally isomorphic to the normalization of the hyperelliptic curve $$w^2=z^5-(a^2+a^{-2})z^3+z, $$ for a unique value of $a\\in(0,1)$ which is approximately $0.0983562$. It is also conformally isomorphic to the quotient of the unit disc by a certain Fuchsian group which again depends on a single parameter $b\\in(0,1)$ which is approximately $0.8005319$. There is a very long story behind all this, which is spelled out in my memoir \"Uniformization of embedded surfaces\" at <a href=\"https://arxiv.org/abs/1607.06433\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1607.06433</a>. There is also a large body of associated Maple code and pictures which can be downloaded from the arxiv, or more conveniently from <a href=\"https://neilstrickland.github.io/genus2/\" rel=\"nofollow noreferrer\">https://neilstrickland.github.io/genus2/</a>.</p>\n" }, { "answer_id": 259603, "score": 4, "text": "<p>Taking this as two questions, the second being more interesting than the first, I can at least answer the first (less interesting) question. The answer is 'Yes, such maps $\\phi:\\mathbb{C}^2\\to\\mathbb{R}^3$ do exist, though they (necessarily) depend on the smooth curve $X$'.</p>\n\n<p>Given a smooth algebraic curve $X\\subset\\mathbb{C}^2$, suppose that $X$ is defined as the zero locus of an reduced polynomial $F(z,w)$ (i.e., $F$ has no multiple factors). Then, by the assumption that $X$ is smooth (by which I assume that David means 'smooth, embedded'), the polynomials, $F$, $F_z$ and $F_w$ have no common zeros. In fact, for each $x\\in X$, the vector $N(x)=\\bigl(\\,\\overline{F_z(x)},\\,\\overline{F_w(x)}\\,\\bigr)$ is nonzero and (unitarily) orthogonal to the (complex) tangent line to $X$ at $x$, since the $1$-form $\\mathrm{d}F = F_z\\,\\mathrm{d}z+F_w\\,\\mathrm{d}w$ vanishes when pulled back to the curve $X$. Let $U(x)= N(x)/|N(x)|$ be the corresponding unit vector.</p>\n\n<p>Now, because $X$ is smooth and algebraic, outside a compact set, it is asymptotic to a finite set of lines, and it is not difficult to see that there is a positive function $e:X\\to (0,1)$ such that the mapping $S:X\\times\\Delta(1)\\to\\mathbb{C}^2$ (where $\\Delta(r)\\subset\\mathbb{C}$ is the disk of radius $r&gt;0$ about $0$) defined by\n$$\nS(x,t) = x + t\\,e(x)U(x)\n$$\nis an injective diffeomorphism. (If all of the asmptotic lines of $X$ are distinct, one can even take $e$ to be a (suitably small) constant.)</p>\n\n<p>Now, we also know that $X$ is a compact (oriented) Riemann surface with a (nonzero) finite number of points removed. (In fact, $X$ has no compact components.) As such, there exists a smooth, closed embedding $\\psi: X\\to \\mathbb{R}^3$ with the property that the normal 'tube' of radius $1$ around $\\psi(X)$ is also smoothly embedded. Let $u:X\\to S^2$ be a unit normal vector field for the immersion $\\psi$ and extend $\\psi$ to $\\psi:X\\times [-\\tfrac12,\\tfrac12]\\to\\mathbb{R}^3$ by setting\n$$\n\\psi(x,t) = \\psi(x) + t\\,u(x)\n$$\nfor $|t|\\le \\tfrac12$. Now, I claim that there is a smooth map $\\phi:S\\bigl(X\\times \\Delta(\\tfrac12)\\bigr)\\to \\mathbb{R}^3$ that satisfies\n$$\n\\phi\\bigl(S(x,t)\\bigr) = \\psi\\bigl(x,\\mathrm{Re}(t)\\bigr)\n$$\nwhenever $|t|\\le\\tfrac12$ and that it is a smooth submersion on an open neighborhood of $X\\subset S\\bigl(X\\times \\Delta(\\tfrac12)\\bigr)$ that is injective and immersive on $X$ itself.</p>\n\n<p>Now, extend $\\phi$ smoothly any way one likes beyond the set $S\\bigl(X\\times \\Delta(\\tfrac12)\\bigr)\\subset\\mathbb{C}^2$. (One can even require that $\\phi$ take the complement of $S\\bigl(X\\times \\Delta(1)\\bigr)$ to a single point of $\\mathbb{R}^3$.)</p>\n\n<p>Unfortunately, there may be no simple recipe for choosing $\\psi$, which is what one would really need to get an affirmative answer to the second question.</p>\n" }, { "answer_id": 259607, "score": 1, "text": "<p>Henry Segerman has asked a similar <a href=\"https://mathoverflow.net/questions/200172/map-of-the-klein-quartic-from-cp2-to-r3\">question</a> about curves in $\\mathbb{CP}^2$ and then managed to produce the <a href=\"http://gallery.bridgesmathart.org/exhibitions/2016-joint-mathematics-meetings/schleimer-segerman\" rel=\"nofollow noreferrer\">sculpture</a> (of the Klein quartic). So perhaps he will have something interesting to say...</p>\n" }, { "answer_id": 260495, "score": 1, "text": "<p>Consider the surface in $\\mathbb{R}^4$ with equations \n\\begin{align*}\n x_1^2+x_2^2+x_3^2+x_4^2 &amp;=1 \\\\\n(3x_3^2-2)x_4-\\sqrt{2}(x_1^2-x_2^2)x_3 &amp;=0.\n\\end{align*} \n<a href=\"https://i.stack.imgur.com/4E9WV.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/4E9WV.png\" alt=\"enter image description here\"></a></p>\n\n<p>This is conformally isomorphic to the normalization of the hyperelliptic curve $$w^2=z^5-(a^2+a^{-2})z^3+z, $$ for a unique value of $a\\in(0,1)$ which is approximately $0.0983562$. It is also conformally isomorphic to the quotient of the unit disc by a certain Fuchsian group which again depends on a single parameter $b\\in(0,1)$ which is approximately $0.8005319$. There is a very long story behind all this, which is spelled out in my memoir \"Uniformization of embedded surfaces\" at <a href=\"https://arxiv.org/abs/1607.06433\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1607.06433</a>. There is also a large body of associated Maple code and pictures which can be downloaded from the arxiv, or more conveniently from <a href=\"https://neilstrickland.github.io/genus2/\" rel=\"nofollow noreferrer\">https://neilstrickland.github.io/genus2/</a>.</p>\n" }, { "answer_id": 259603, "score": 4, "text": "<p>Taking this as two questions, the second being more interesting than the first, I can at least answer the first (less interesting) question. The answer is 'Yes, such maps $\\phi:\\mathbb{C}^2\\to\\mathbb{R}^3$ do exist, though they (necessarily) depend on the smooth curve $X$'.</p>\n\n<p>Given a smooth algebraic curve $X\\subset\\mathbb{C}^2$, suppose that $X$ is defined as the zero locus of an reduced polynomial $F(z,w)$ (i.e., $F$ has no multiple factors). Then, by the assumption that $X$ is smooth (by which I assume that David means 'smooth, embedded'), the polynomials, $F$, $F_z$ and $F_w$ have no common zeros. In fact, for each $x\\in X$, the vector $N(x)=\\bigl(\\,\\overline{F_z(x)},\\,\\overline{F_w(x)}\\,\\bigr)$ is nonzero and (unitarily) orthogonal to the (complex) tangent line to $X$ at $x$, since the $1$-form $\\mathrm{d}F = F_z\\,\\mathrm{d}z+F_w\\,\\mathrm{d}w$ vanishes when pulled back to the curve $X$. Let $U(x)= N(x)/|N(x)|$ be the corresponding unit vector.</p>\n\n<p>Now, because $X$ is smooth and algebraic, outside a compact set, it is asymptotic to a finite set of lines, and it is not difficult to see that there is a positive function $e:X\\to (0,1)$ such that the mapping $S:X\\times\\Delta(1)\\to\\mathbb{C}^2$ (where $\\Delta(r)\\subset\\mathbb{C}$ is the disk of radius $r&gt;0$ about $0$) defined by\n$$\nS(x,t) = x + t\\,e(x)U(x)\n$$\nis an injective diffeomorphism. (If all of the asmptotic lines of $X$ are distinct, one can even take $e$ to be a (suitably small) constant.)</p>\n\n<p>Now, we also know that $X$ is a compact (oriented) Riemann surface with a (nonzero) finite number of points removed. (In fact, $X$ has no compact components.) As such, there exists a smooth, closed embedding $\\psi: X\\to \\mathbb{R}^3$ with the property that the normal 'tube' of radius $1$ around $\\psi(X)$ is also smoothly embedded. Let $u:X\\to S^2$ be a unit normal vector field for the immersion $\\psi$ and extend $\\psi$ to $\\psi:X\\times [-\\tfrac12,\\tfrac12]\\to\\mathbb{R}^3$ by setting\n$$\n\\psi(x,t) = \\psi(x) + t\\,u(x)\n$$\nfor $|t|\\le \\tfrac12$. Now, I claim that there is a smooth map $\\phi:S\\bigl(X\\times \\Delta(\\tfrac12)\\bigr)\\to \\mathbb{R}^3$ that satisfies\n$$\n\\phi\\bigl(S(x,t)\\bigr) = \\psi\\bigl(x,\\mathrm{Re}(t)\\bigr)\n$$\nwhenever $|t|\\le\\tfrac12$ and that it is a smooth submersion on an open neighborhood of $X\\subset S\\bigl(X\\times \\Delta(\\tfrac12)\\bigr)$ that is injective and immersive on $X$ itself.</p>\n\n<p>Now, extend $\\phi$ smoothly any way one likes beyond the set $S\\bigl(X\\times \\Delta(\\tfrac12)\\bigr)\\subset\\mathbb{C}^2$. (One can even require that $\\phi$ take the complement of $S\\bigl(X\\times \\Delta(1)\\bigr)$ to a single point of $\\mathbb{R}^3$.)</p>\n\n<p>Unfortunately, there may be no simple recipe for choosing $\\psi$, which is what one would really need to get an affirmative answer to the second question.</p>\n" }, { "answer_id": 259607, "score": 1, "text": "<p>Henry Segerman has asked a similar <a href=\"https://mathoverflow.net/questions/200172/map-of-the-klein-quartic-from-cp2-to-r3\">question</a> about curves in $\\mathbb{CP}^2$ and then managed to produce the <a href=\"http://gallery.bridgesmathart.org/exhibitions/2016-joint-mathematics-meetings/schleimer-segerman\" rel=\"nofollow noreferrer\">sculpture</a> (of the Klein quartic). So perhaps he will have something interesting to say...</p>\n" }, { "answer_id": 260495, "score": 1, "text": "<p>Consider the surface in $\\mathbb{R}^4$ with equations \n\\begin{align*}\n x_1^2+x_2^2+x_3^2+x_4^2 &amp;=1 \\\\\n(3x_3^2-2)x_4-\\sqrt{2}(x_1^2-x_2^2)x_3 &amp;=0.\n\\end{align*} \n<a href=\"https://i.stack.imgur.com/4E9WV.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/4E9WV.png\" alt=\"enter image description here\"></a></p>\n\n<p>This is conformally isomorphic to the normalization of the hyperelliptic curve $$w^2=z^5-(a^2+a^{-2})z^3+z, $$ for a unique value of $a\\in(0,1)$ which is approximately $0.0983562$. It is also conformally isomorphic to the quotient of the unit disc by a certain Fuchsian group which again depends on a single parameter $b\\in(0,1)$ which is approximately $0.8005319$. There is a very long story behind all this, which is spelled out in my memoir \"Uniformization of embedded surfaces\" at <a href=\"https://arxiv.org/abs/1607.06433\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1607.06433</a>. There is also a large body of associated Maple code and pictures which can be downloaded from the arxiv, or more conveniently from <a href=\"https://neilstrickland.github.io/genus2/\" rel=\"nofollow noreferrer\">https://neilstrickland.github.io/genus2/</a>.</p>\n" } ]
Question: Putting algebraic curves in $\mathbb{R}^3$ Answer 1: <p>Taking this as two questions, the second being more interesting than the first, I can at least answer the first (less interesting) question. The answer is 'Yes, such maps $\phi:\mathbb{C}^2\to\mathbb{R}^3$ do exist, though they (necessarily) depend on the smooth curve $X$'.</p> <p>Given a smooth algebraic curve $X\subset\mathbb{C}^2$, suppose that $X$ is defined as the zero locus of an reduced polynomial $F(z,w)$ (i.e., $F$ has no multiple factors). Then, by the assumption that $X$ is smooth (by which I assume that David means 'smooth, embedded'), the polynomials, $F$, $F_z$ and $F_w$ have no common zeros. In fact, for each $x\in X$, the vector $N(x)=\bigl(\,\overline{F_z(x)},\,\overline{F_w(x)}\,\bigr)$ is nonzero and (unitarily) orthogonal to the (complex) tangent line to $X$ at $x$, since the $1$-form $\mathrm{d}F = F_z\,\mathrm{d}z+F_w\,\mathrm{d}w$ vanishes when pulled back to the curve $X$. Let $U(x)= N(x)/|N(x)|$ be the corresponding unit vector.</p> <p>Now, because $X$ is smooth and algebraic, outside a compact set, it is asymptotic to a finite set of lines, and it is not difficult to see that there is a positive function $e:X\to (0,1)$ such that the mapping $S:X\times\Delta(1)\to\mathbb{C}^2$ (where $\Delta(r)\subset\mathbb{C}$ is the disk of radius $r&gt;0$ about $0$) defined by $$ S(x,t) = x + t\,e(x)U(x) $$ is an injective diffeomorphism. (If all of the asmptotic lines of $X$ are distinct, one can even take $e$ to be a (suitably small) constant.)</p> <p>Now, we also know that $X$ is a compact (oriented) Riemann surface with a (nonzero) finite number of points removed. (In fact, $X$ has no compact components.) As such, there exists a smooth, closed embedding $\psi: X\to \mathbb{R}^3$ with the property that the normal 'tube' of radius $1$ around $\psi(X)$ is also smoothly embedded. Let $u:X\to S^2$ be a unit normal vector field for the immersion $\psi$ and extend $\psi$ to $\psi:X\times [-\tfrac12,\tfrac12]\to\mathbb{R}^3$ by setting $$ \psi(x,t) = \psi(x) + t\,u(x) $$ for $|t|\le \tfrac12$. Now, I claim that there is a smooth map $\phi:S\bigl(X\times \Delta(\tfrac12)\bigr)\to \mathbb{R}^3$ that satisfies $$ \phi\bigl(S(x,t)\bigr) = \psi\bigl(x,\mathrm{Re}(t)\bigr) $$ whenever $|t|\le\tfrac12$ and that it is a smooth submersion on an open neighborhood of $X\subset S\bigl(X\times \Delta(\tfrac12)\bigr)$ that is injective and immersive on $X$ itself.</p> <p>Now, extend $\phi$ smoothly any way one likes beyond the set $S\bigl(X\times \Delta(\tfrac12)\bigr)\subset\mathbb{C}^2$. (One can even require that $\phi$ take the complement of $S\bigl(X\times \Delta(1)\bigr)$ to a single point of $\mathbb{R}^3$.)</p> <p>Unfortunately, there may be no simple recipe for choosing $\psi$, which is what one would really need to get an affirmative answer to the second question.</p> Answer 2: <p>Henry Segerman has asked a similar <a href="https://mathoverflow.net/questions/200172/map-of-the-klein-quartic-from-cp2-to-r3">question</a> about curves in $\mathbb{CP}^2$ and then managed to produce the <a href="http://gallery.bridgesmathart.org/exhibitions/2016-joint-mathematics-meetings/schleimer-segerman" rel="nofollow noreferrer">sculpture</a> (of the Klein quartic). So perhaps he will have something interesting to say...</p> Answer 3: <p>Consider the surface in $\mathbb{R}^4$ with equations \begin{align*} x_1^2+x_2^2+x_3^2+x_4^2 &amp;=1 \\ (3x_3^2-2)x_4-\sqrt{2}(x_1^2-x_2^2)x_3 &amp;=0. \end{align*} <a href="https://i.stack.imgur.com/4E9WV.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/4E9WV.png" alt="enter image description here"></a></p> <p>This is conformally isomorphic to the normalization of the hyperelliptic curve $$w^2=z^5-(a^2+a^{-2})z^3+z, $$ for a unique value of $a\in(0,1)$ which is approximately $0.0983562$. It is also conformally isomorphic to the quotient of the unit disc by a certain Fuchsian group which again depends on a single parameter $b\in(0,1)$ which is approximately $0.8005319$. There is a very long story behind all this, which is spelled out in my memoir "Uniformization of embedded surfaces" at <a href="https://arxiv.org/abs/1607.06433" rel="nofollow noreferrer">https://arxiv.org/abs/1607.06433</a>. There is also a large body of associated Maple code and pictures which can be downloaded from the arxiv, or more conveniently from <a href="https://neilstrickland.github.io/genus2/" rel="nofollow noreferrer">https://neilstrickland.github.io/genus2/</a>.</p> Answer 4: <p>Taking this as two questions, the second being more interesting than the first, I can at least answer the first (less interesting) question. The answer is 'Yes, such maps $\phi:\mathbb{C}^2\to\mathbb{R}^3$ do exist, though they (necessarily) depend on the smooth curve $X$'.</p> <p>Given a smooth algebraic curve $X\subset\mathbb{C}^2$, suppose that $X$ is defined as the zero locus of an reduced polynomial $F(z,w)$ (i.e., $F$ has no multiple factors). Then, by the assumption that $X$ is smooth (by which I assume that David means 'smooth, embedded'), the polynomials, $F$, $F_z$ and $F_w$ have no common zeros. In fact, for each $x\in X$, the vector $N(x)=\bigl(\,\overline{F_z(x)},\,\overline{F_w(x)}\,\bigr)$ is nonzero and (unitarily) orthogonal to the (complex) tangent line to $X$ at $x$, since the $1$-form $\mathrm{d}F = F_z\,\mathrm{d}z+F_w\,\mathrm{d}w$ vanishes when pulled back to the curve $X$. Let $U(x)= N(x)/|N(x)|$ be the corresponding unit vector.</p> <p>Now, because $X$ is smooth and algebraic, outside a compact set, it is asymptotic to a finite set of lines, and it is not difficult to see that there is a positive function $e:X\to (0,1)$ such that the mapping $S:X\times\Delta(1)\to\mathbb{C}^2$ (where $\Delta(r)\subset\mathbb{C}$ is the disk of radius $r&gt;0$ about $0$) defined by $$ S(x,t) = x + t\,e(x)U(x) $$ is an injective diffeomorphism. (If all of the asmptotic lines of $X$ are distinct, one can even take $e$ to be a (suitably small) constant.)</p> <p>Now, we also know that $X$ is a compact (oriented) Riemann surface with a (nonzero) finite number of points removed. (In fact, $X$ has no compact components.) As such, there exists a smooth, closed embedding $\psi: X\to \mathbb{R}^3$ with the property that the normal 'tube' of radius $1$ around $\psi(X)$ is also smoothly embedded. Let $u:X\to S^2$ be a unit normal vector field for the immersion $\psi$ and extend $\psi$ to $\psi:X\times [-\tfrac12,\tfrac12]\to\mathbb{R}^3$ by setting $$ \psi(x,t) = \psi(x) + t\,u(x) $$ for $|t|\le \tfrac12$. Now, I claim that there is a smooth map $\phi:S\bigl(X\times \Delta(\tfrac12)\bigr)\to \mathbb{R}^3$ that satisfies $$ \phi\bigl(S(x,t)\bigr) = \psi\bigl(x,\mathrm{Re}(t)\bigr) $$ whenever $|t|\le\tfrac12$ and that it is a smooth submersion on an open neighborhood of $X\subset S\bigl(X\times \Delta(\tfrac12)\bigr)$ that is injective and immersive on $X$ itself.</p> <p>Now, extend $\phi$ smoothly any way one likes beyond the set $S\bigl(X\times \Delta(\tfrac12)\bigr)\subset\mathbb{C}^2$. (One can even require that $\phi$ take the complement of $S\bigl(X\times \Delta(1)\bigr)$ to a single point of $\mathbb{R}^3$.)</p> <p>Unfortunately, there may be no simple recipe for choosing $\psi$, which is what one would really need to get an affirmative answer to the second question.</p> Answer 5: <p>Henry Segerman has asked a similar <a href="https://mathoverflow.net/questions/200172/map-of-the-klein-quartic-from-cp2-to-r3">question</a> about curves in $\mathbb{CP}^2$ and then managed to produce the <a href="http://gallery.bridgesmathart.org/exhibitions/2016-joint-mathematics-meetings/schleimer-segerman" rel="nofollow noreferrer">sculpture</a> (of the Klein quartic). So perhaps he will have something interesting to say...</p> Answer 6: <p>Consider the surface in $\mathbb{R}^4$ with equations \begin{align*} x_1^2+x_2^2+x_3^2+x_4^2 &amp;=1 \\ (3x_3^2-2)x_4-\sqrt{2}(x_1^2-x_2^2)x_3 &amp;=0. \end{align*} <a href="https://i.stack.imgur.com/4E9WV.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/4E9WV.png" alt="enter image description here"></a></p> <p>This is conformally isomorphic to the normalization of the hyperelliptic curve $$w^2=z^5-(a^2+a^{-2})z^3+z, $$ for a unique value of $a\in(0,1)$ which is approximately $0.0983562$. It is also conformally isomorphic to the quotient of the unit disc by a certain Fuchsian group which again depends on a single parameter $b\in(0,1)$ which is approximately $0.8005319$. There is a very long story behind all this, which is spelled out in my memoir "Uniformization of embedded surfaces" at <a href="https://arxiv.org/abs/1607.06433" rel="nofollow noreferrer">https://arxiv.org/abs/1607.06433</a>. There is also a large body of associated Maple code and pictures which can be downloaded from the arxiv, or more conveniently from <a href="https://neilstrickland.github.io/genus2/" rel="nofollow noreferrer">https://neilstrickland.github.io/genus2/</a>.</p> Answer 7: <p>Taking this as two questions, the second being more interesting than the first, I can at least answer the first (less interesting) question. The answer is 'Yes, such maps $\phi:\mathbb{C}^2\to\mathbb{R}^3$ do exist, though they (necessarily) depend on the smooth curve $X$'.</p> <p>Given a smooth algebraic curve $X\subset\mathbb{C}^2$, suppose that $X$ is defined as the zero locus of an reduced polynomial $F(z,w)$ (i.e., $F$ has no multiple factors). Then, by the assumption that $X$ is smooth (by which I assume that David means 'smooth, embedded'), the polynomials, $F$, $F_z$ and $F_w$ have no common zeros. In fact, for each $x\in X$, the vector $N(x)=\bigl(\,\overline{F_z(x)},\,\overline{F_w(x)}\,\bigr)$ is nonzero and (unitarily) orthogonal to the (complex) tangent line to $X$ at $x$, since the $1$-form $\mathrm{d}F = F_z\,\mathrm{d}z+F_w\,\mathrm{d}w$ vanishes when pulled back to the curve $X$. Let $U(x)= N(x)/|N(x)|$ be the corresponding unit vector.</p> <p>Now, because $X$ is smooth and algebraic, outside a compact set, it is asymptotic to a finite set of lines, and it is not difficult to see that there is a positive function $e:X\to (0,1)$ such that the mapping $S:X\times\Delta(1)\to\mathbb{C}^2$ (where $\Delta(r)\subset\mathbb{C}$ is the disk of radius $r&gt;0$ about $0$) defined by $$ S(x,t) = x + t\,e(x)U(x) $$ is an injective diffeomorphism. (If all of the asmptotic lines of $X$ are distinct, one can even take $e$ to be a (suitably small) constant.)</p> <p>Now, we also know that $X$ is a compact (oriented) Riemann surface with a (nonzero) finite number of points removed. (In fact, $X$ has no compact components.) As such, there exists a smooth, closed embedding $\psi: X\to \mathbb{R}^3$ with the property that the normal 'tube' of radius $1$ around $\psi(X)$ is also smoothly embedded. Let $u:X\to S^2$ be a unit normal vector field for the immersion $\psi$ and extend $\psi$ to $\psi:X\times [-\tfrac12,\tfrac12]\to\mathbb{R}^3$ by setting $$ \psi(x,t) = \psi(x) + t\,u(x) $$ for $|t|\le \tfrac12$. Now, I claim that there is a smooth map $\phi:S\bigl(X\times \Delta(\tfrac12)\bigr)\to \mathbb{R}^3$ that satisfies $$ \phi\bigl(S(x,t)\bigr) = \psi\bigl(x,\mathrm{Re}(t)\bigr) $$ whenever $|t|\le\tfrac12$ and that it is a smooth submersion on an open neighborhood of $X\subset S\bigl(X\times \Delta(\tfrac12)\bigr)$ that is injective and immersive on $X$ itself.</p> <p>Now, extend $\phi$ smoothly any way one likes beyond the set $S\bigl(X\times \Delta(\tfrac12)\bigr)\subset\mathbb{C}^2$. (One can even require that $\phi$ take the complement of $S\bigl(X\times \Delta(1)\bigr)$ to a single point of $\mathbb{R}^3$.)</p> <p>Unfortunately, there may be no simple recipe for choosing $\psi$, which is what one would really need to get an affirmative answer to the second question.</p> Answer 8: <p>Henry Segerman has asked a similar <a href="https://mathoverflow.net/questions/200172/map-of-the-klein-quartic-from-cp2-to-r3">question</a> about curves in $\mathbb{CP}^2$ and then managed to produce the <a href="http://gallery.bridgesmathart.org/exhibitions/2016-joint-mathematics-meetings/schleimer-segerman" rel="nofollow noreferrer">sculpture</a> (of the Klein quartic). So perhaps he will have something interesting to say...</p> Answer 9: <p>Consider the surface in $\mathbb{R}^4$ with equations \begin{align*} x_1^2+x_2^2+x_3^2+x_4^2 &amp;=1 \\ (3x_3^2-2)x_4-\sqrt{2}(x_1^2-x_2^2)x_3 &amp;=0. \end{align*} <a href="https://i.stack.imgur.com/4E9WV.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/4E9WV.png" alt="enter image description here"></a></p> <p>This is conformally isomorphic to the normalization of the hyperelliptic curve $$w^2=z^5-(a^2+a^{-2})z^3+z, $$ for a unique value of $a\in(0,1)$ which is approximately $0.0983562$. It is also conformally isomorphic to the quotient of the unit disc by a certain Fuchsian group which again depends on a single parameter $b\in(0,1)$ which is approximately $0.8005319$. There is a very long story behind all this, which is spelled out in my memoir "Uniformization of embedded surfaces" at <a href="https://arxiv.org/abs/1607.06433" rel="nofollow noreferrer">https://arxiv.org/abs/1607.06433</a>. There is also a large body of associated Maple code and pictures which can be downloaded from the arxiv, or more conveniently from <a href="https://neilstrickland.github.io/genus2/" rel="nofollow noreferrer">https://neilstrickland.github.io/genus2/</a>.</p>
A seemingly simple inequality
259,637
18
https://mathoverflow.net/questions/259637/a-seemingly-simple-inequality
<p>Let $a_i,b_i\in\mathbb{R}$ and $n&gt;1$, does the inequality $$ \left(\sum_{i=1}^n a_i^2\right)\left(\sum_{i=1}^n b_i^2\right)+\left(\sum_{i=1}^na_i b_i\right)^2\ge \sqrt{\left(\sum_{i=1}^n a_i^4\right)\left(\sum_{i=1}^n b_i^4\right)}+\sum_{i=1}^na_i^2b_i^2 $$ hold true?</p> <p>Observe that $a_i$ and $b_i$ are <strong>not</strong> required to be non-negative. I ran an extensive number of numerical simulations and no counterexample showed up yet.</p> <p><strong>Note 1.</strong> The inequality holds true for $n=2$, as showed <a href="http://www.wolframalpha.com/input/?i=((a_1%5E2%2Ba_2%5E2)(b_1%5E2%2Bb_2%5E2)%2B(a_1*b_1%2Ba_2*b_2)%5E2-(a_1%5E2*b_1%5E2%2Ba_2%5E2*b_2%5E2))%5E2-(a_1%5E4%2Ba_2%5E4)(b_1%5E4%2Bb_2%5E4)" rel="nofollow noreferrer">here</a>.</p> <p><strong>Note 2.</strong> This conjecture was formulated by Fedor Petrov in an attempt to provide a solution to a particular case of <a href="https://mathoverflow.net/questions/233546/a-log-inequality-for-positive-definite-trace-one-matrices">this question</a>.</p> <p><strong>Note 3.</strong> I have posted this question on <a href="https://math.stackexchange.com/questions/1706416/sum-limits-i-a-i2-sum-limits-i-b-i2-left-sum-limits-ia-i-b-i-right2-geq">math.SE</a> some time ago but it has received no answer, so I cross-posted it here.</p> <p><strong>Note 4.</strong> As Fedor Petrov rightly observed in his answer below, the inequality also follows by an argument used in one of his answers in the above-cited question. However, I decided to accept Markus Sprecher's answer because of its conciseness and clarity.</p>
[ { "answer_id": 259675, "score": 19, "text": "<p>The inequality is equivalent to\n$$\n\\left(\\sum_{i&gt;j} (a_ib_j+a_jb_i)^2+\\sum_{i} a^2_ib^2_i \\right)^2\\geq \\sum_{i} a^4_i \\sum_{i} b^4_i.\n$$\nThe left hand side is greater or equal to\n$$\n\\sum_i a_i^4b_i^4+\\sum_{i&gt;j} (a_ib_j+a_jb_i)^4+2(a_ib_j+a_jb_i)^2(a_i^2b_i^2+a_j^2b_j^2)+2a_i^2b_i^2a_j^2b_j^2\n$$\nAs\n$$\n(a_ib_j+a_jb_i)^4+2(a_ib_j+a_jb_i)^2(a_i^2b_i^2+a_j^2b_j^2)+2a_i^2b_i^2a_j^2b_j^2\\geq a_i^4b_j^4+a_j^4b_i^4.\n$$\nis equivalent to\n$$\n(a_ib_j+a_jb_i)^2(a_ib_i+a_jb_j)^2\\geq 0\n$$\nthe LHS is larger or equal to\n$$\n\\sum_i a_i^4b_i^4+\\sum_{i&gt;j} a_i^4b_j^4+a_j^4b_i^4=\\sum_{i} a^4_i \\sum_{i} b^4_i.\n$$</p>\n" }, { "answer_id": 259684, "score": 7, "text": "<p>I have an other solution, which not as slick Markus Sprecher's one, but I think the method, which is quite general, is interesting in itself.</p>\n\n<p>Fix $c \\in \\left]-1,1\\right[$ and $n \\geq 2$, and consider the compact manifold\n$$\n\\Sigma = \\{ (a,b) \\in S^{n-1} \\times S^{n-1} \\ | \\ \\langle a , b \\rangle = c \\}.\n$$\nIt has dimension $2n - 3$, and the tangent space at $(a,b)$ is the orthogonal in $\\mathbb{R}^{2n}$ to the 3-space spanned by $(a,0),(0,b),(b,a)$.</p>\n\n<p>Consider the functions\n$$\nr_1 = \\sqrt{\\sum_i a_i^4}, \\ \\ r_2 = \\sqrt{\\sum_i b_i^4}, \\ \\ r_3 = \\sum_{i} a_i^2 b_i^2\n$$\nas continuous functions on $\\Sigma$. Let $(a,b)$ be a point of $\\Sigma$ where $f = r_1 r_2 + r_3$ is maximal. We would like to show that $f(a,b) \\leq 1 + c^2$.</p>\n\n<p>The gradient of $f$ at $(a,b)$ must be of the form $\\alpha(a,0) + \\beta(0,b) + \\gamma(b,a)$ for some real numbers $\\alpha,\\beta,\\gamma$. This gives the following $2n$ equations:\n$$\n(1) \\ \\ a_i b_i^2 + a_i^3 \\frac{r_2}{r_1} = \\alpha a_i + \\gamma b_i \\\\\n(2) \\ \\ b_i a_i^2 + b_i^3 \\frac{r_1}{r_2} = \\beta b_i + \\gamma a_i \\\\\n$$\nMultiplying $(1)$ by $a_i$ and summing ober $i$, we get $f = r_3 +r_1r_2 = \\alpha + c \\gamma$. \nMultiplying $(2)$ by $b_i$ and summing ober $i$, we get $f = r_3 +r_1r_2 = \\beta + c \\gamma$. In particular $\\alpha = \\beta$.</p>\n\n<p>Mutliply equations $(1)$ and $(2)$ by $r_1 b_i$ and $r_2 a_i$ respectively. This yields\n$$\n(1)' \\ \\ r_1 a_i b_i^3 + r_2 a_i^3 b_i = r_1 \\alpha a_i b_i + r_1 \\gamma b_i^2 \\\\\n(2)' \\ \\ r_2 a_i^3 b_i + r_1 a_i b_i^3 = r_2 \\alpha a_ib_i + r_2 \\gamma a_i^2 \\\\\n$$\nThe two LHSs are equal, hence so are the RHSs. Thus $(a_i,b_i)$ satisifies the quadratic equation\n$$\n(3) r_1 \\alpha a_i b_i + r_1 \\gamma b_i^2 = r_2 \\alpha a_ib_i + r_2 \\gamma a_i^2.\n$$</p>\n\n<p>There are two cases:</p>\n\n<ul>\n<li>if $\\gamma =0$, then $\\alpha = f &gt;0$, and thus $r_1 = r_2$ follows from $(3)$. Then $(1)-(2)$ become the alternative $(a_i,b_i) =0$ or $a_i^2+ b_i^2 = \\alpha$. Since $a$ and $b$ are not proportional, the latter cas happens at least twice, and thus $2 = ||a||_2^2 + ||b||_2^2 \\geq 2 \\alpha$. Thus $f = \\alpha \\leq 1$, and in particular $f \\leq 1 + c^2$.</li>\n<li>if $\\gamma \\neq 0$ then there are at most $2$ projective solutions to $(3)$. Since the system $(1)-(2)$ is inhomogeneous, it has at most $2$ non-zero solutions up to sign. Thus there are non zero vectors $(a_1'',b_1'')$ and $(a_2'',b_2'')$ such that for eachi $i$ one has either $(a_i,b_i) = \\pm (a_j'',b_j'')$ for some $j=1,2$, or $(a_i,b_i) = 0$. Let $[|1,n|] = I_0 \\sqcup I_1 \\sqcup I_2$ such that $(a_i,b_i) = 0$ for $i$ in $I_0$, and $(a_i,b_i) = \\pm (a_j'',b_j'')$ if $i$ is in $I_j$ for some $j=1,2$. We then have the identities\n\\begin{align*}\n|I_1| (a_1'')^2 + |I_2| (a_2'')^2 &amp;= \\sum_i a_i^2 = 1 \\\\\n|I_1| a_1''b_1'' + |I_2| a_2'' b_2'' &amp;= \\sum_i a_i b_i = c \\\\\n|I_1| (b_1'')^2 + |I_2| (b_2'')^2 &amp;= \\sum_i b_i^2 = 1 \n\\end{align*}\nThus the vectors $a' = (|I_1|^{\\frac{1}{2}} a_1'',|I_2|^{\\frac{1}{2}}a_2'')$ and $b' = (|I_1|^{\\frac{1}{2}} b_1'',|I_2|^{\\frac{1}{2}}b_2'')$ in $\\mathbb{R}^2$ satisfy\n$$\n||a'||_2 =1, \\ \\ ||b'|| = 1, \\ \\ \\langle a',b' \\rangle = c.\n$$\nMoreover, one has\n\\begin{align*}\n\\sum_{i=1}^2 (a_i')^4 &amp;= |I_1|^2 (a_1'')^4 + |I_2|^2 (a_2'')^4 \\geq |I_1| (a_1'')^4 + |I_2| (a_2'')^4 = r_1^2 \\\\\n\\sum_{i=1}^2 (b_i')^4 &amp;= |I_1|^2 (b_1'')^4 + |I_2|^2 (b_2'')^4 \\geq |I_1| (b_1'')^4 + |I_2| (b_2'')^4 = r_2^2 \\\\\n\\sum_{i=1}^2 (a_i')^2(b_i')^2 &amp;= |I_1|^2 (a_1'')^2(b_1'')^2 + |I_2|^2 (a_2'')^2(b_2'')^2 \\geq |I_1| (a_1'')^2(b_1'')^2 + |I_2| (a_2'')^2 (b_2'')^2 = r_3 \\\\\n\\end{align*}\nConsequently, applying the case $n=2$ of the inequality to the vectors $a',b'$ yields\n$$\n1 + c^2 \\geq r_1 r_2 + r_3 = f(a,b).\n$$\nThus $f \\leq 1+c^2$ on all of $\\Sigma$.</li>\n</ul>\n" }, { "answer_id": 259934, "score": 5, "text": "<p>Just to clarify. As I remember, this is simply equivalent to the (partial case of) the question you cite. Since the cited question was solved, I would not call it a conjecture. But the question to find an independent proof makes sense. </p>\n\n<p>Well, let me elaborate. <a href=\"https://mathoverflow.net/q/234218\">Here</a> proving the partial case $N=2$ of your inequality I prove the following inequality: $$\\frac{\\|x\\|^2\\cdot \\|y\\|^2+(x,y)^2}{\\|Tx\\|^2 \\|Ty\\|^2}\\geqslant \\frac2{{\\rm tr}\\, T^4}$$\nfor any self-adjoint positive definite operator $T$ and any two vectors $x,y$ in $\\mathbb{R}^n$. If we denote $x=(a_1,\\dots,a_n)$, $y=(b_1,\\dots,b_n)$, $p=(a_1^2,\\dots,a_n^2)$, $q=(b_1^2,\\dots,b_n^2)$, $T^2=diag(s_1,\\dots,s_n)$, $s=(s_1,\\dots,s_n)$, we rewrite this as $$\\left(\\sum_{i=1}^n a_i^2\\right)\\left(\\sum_{i=1}^n b_i^2\\right)+\\left(\\sum_{i=1}^na_i b_i\\right)^2\\ge 2\\frac{(p,s)\\cdot (q,s)}{(s,s)},$$\nand maximizing over $s$ gives you $$\\sqrt{\\left(\\sum_{i=1}^n a_i^4\\right)\\left(\\sum_{i=1}^n b_i^4\\right)}+\\sum_{i=1}^na_i^2b_i^2$$ in RHS, this is the equality case in the lemma in the cited answer.</p>\n" }, { "answer_id": 259699, "score": 2, "text": "<p>Denote the power-sum (symmetric) polynomials by $p_k(a;n)=\\sum_{i=1}^na_i^k$; where $a=(a_1,\\dots, a_n)$ and also we need the Hadamard product $ab=(a_1b_1,\\dots,a_nb_n)$. Write $p_k(a)$ for $p_k(a;n)$ when there is no confusion. </p>\n\n<p><strong>Remark.</strong> Observe that $p_2(a)p_2(b)-p_2(ab)=\\sum_{i\\neq j}a_i^2b_j^2\\geq0$. We prove\n$$[p_2(a)p_2(b)+p_1(ab)^2-p_2(ab)]^2-p_2(a^2)p_2(b^2)\\geq0.\\tag1$$\nInduct on $n$. The case $n=2$ is noted as trivial. Assume true for $n$, we show for $n+1$; i.e.\n$$[(x^2+p_2(a))(y^2+p_2(b))+(xy+p_1(ab))^2-x^2y^2-p_2(ab)]^2-(x^4+p_2(a^2))(y^4+p_2(b^2))\\geq0.\\tag2$$\nAfter expansion, the LHS of (2) is a polynomial in even powers of $x$ and $y$. It suffices to study the following coefficients (the others follow by symmetry):</p>\n\n<p>$[x^0y^0]$: the is exactly the induction assumption, hence positive.</p>\n\n<p>$[x^2y^0]$: $2p_2(b)[p_1(ab)^2+p_2(a)p_2(b)-p_2(ab)]\\geq0$, by remark from above.</p>\n\n<p>$x^4y^0]$: $p_2(b)^2-p_2(b^2)\\geq0$, by remark.</p>\n\n<p>$[x^2y^2]$: $6p_1(ab)^2+4p_2(a)p_2(b)-2p_2(ab)\\geq0$, by remark.</p>\n\n<p>$[x^4y^2]$: $2p_2(b)\\geq0$.</p>\n\n<p>$[x^0y^4]$: $p_2(a)^2-p_2(a^2)\\geq0$, by remark.</p>\n\n<p>Therefore, the claim (1) is valid for all $n$.</p>\n\n<p>The intent of this method is to reveal that the inequality (1) is not as sharp as we may wish.</p>\n" }, { "answer_id": 259675, "score": 19, "text": "<p>The inequality is equivalent to\n$$\n\\left(\\sum_{i&gt;j} (a_ib_j+a_jb_i)^2+\\sum_{i} a^2_ib^2_i \\right)^2\\geq \\sum_{i} a^4_i \\sum_{i} b^4_i.\n$$\nThe left hand side is greater or equal to\n$$\n\\sum_i a_i^4b_i^4+\\sum_{i&gt;j} (a_ib_j+a_jb_i)^4+2(a_ib_j+a_jb_i)^2(a_i^2b_i^2+a_j^2b_j^2)+2a_i^2b_i^2a_j^2b_j^2\n$$\nAs\n$$\n(a_ib_j+a_jb_i)^4+2(a_ib_j+a_jb_i)^2(a_i^2b_i^2+a_j^2b_j^2)+2a_i^2b_i^2a_j^2b_j^2\\geq a_i^4b_j^4+a_j^4b_i^4.\n$$\nis equivalent to\n$$\n(a_ib_j+a_jb_i)^2(a_ib_i+a_jb_j)^2\\geq 0\n$$\nthe LHS is larger or equal to\n$$\n\\sum_i a_i^4b_i^4+\\sum_{i&gt;j} a_i^4b_j^4+a_j^4b_i^4=\\sum_{i} a^4_i \\sum_{i} b^4_i.\n$$</p>\n" }, { "answer_id": 259684, "score": 7, "text": "<p>I have an other solution, which not as slick Markus Sprecher's one, but I think the method, which is quite general, is interesting in itself.</p>\n\n<p>Fix $c \\in \\left]-1,1\\right[$ and $n \\geq 2$, and consider the compact manifold\n$$\n\\Sigma = \\{ (a,b) \\in S^{n-1} \\times S^{n-1} \\ | \\ \\langle a , b \\rangle = c \\}.\n$$\nIt has dimension $2n - 3$, and the tangent space at $(a,b)$ is the orthogonal in $\\mathbb{R}^{2n}$ to the 3-space spanned by $(a,0),(0,b),(b,a)$.</p>\n\n<p>Consider the functions\n$$\nr_1 = \\sqrt{\\sum_i a_i^4}, \\ \\ r_2 = \\sqrt{\\sum_i b_i^4}, \\ \\ r_3 = \\sum_{i} a_i^2 b_i^2\n$$\nas continuous functions on $\\Sigma$. Let $(a,b)$ be a point of $\\Sigma$ where $f = r_1 r_2 + r_3$ is maximal. We would like to show that $f(a,b) \\leq 1 + c^2$.</p>\n\n<p>The gradient of $f$ at $(a,b)$ must be of the form $\\alpha(a,0) + \\beta(0,b) + \\gamma(b,a)$ for some real numbers $\\alpha,\\beta,\\gamma$. This gives the following $2n$ equations:\n$$\n(1) \\ \\ a_i b_i^2 + a_i^3 \\frac{r_2}{r_1} = \\alpha a_i + \\gamma b_i \\\\\n(2) \\ \\ b_i a_i^2 + b_i^3 \\frac{r_1}{r_2} = \\beta b_i + \\gamma a_i \\\\\n$$\nMultiplying $(1)$ by $a_i$ and summing ober $i$, we get $f = r_3 +r_1r_2 = \\alpha + c \\gamma$. \nMultiplying $(2)$ by $b_i$ and summing ober $i$, we get $f = r_3 +r_1r_2 = \\beta + c \\gamma$. In particular $\\alpha = \\beta$.</p>\n\n<p>Mutliply equations $(1)$ and $(2)$ by $r_1 b_i$ and $r_2 a_i$ respectively. This yields\n$$\n(1)' \\ \\ r_1 a_i b_i^3 + r_2 a_i^3 b_i = r_1 \\alpha a_i b_i + r_1 \\gamma b_i^2 \\\\\n(2)' \\ \\ r_2 a_i^3 b_i + r_1 a_i b_i^3 = r_2 \\alpha a_ib_i + r_2 \\gamma a_i^2 \\\\\n$$\nThe two LHSs are equal, hence so are the RHSs. Thus $(a_i,b_i)$ satisifies the quadratic equation\n$$\n(3) r_1 \\alpha a_i b_i + r_1 \\gamma b_i^2 = r_2 \\alpha a_ib_i + r_2 \\gamma a_i^2.\n$$</p>\n\n<p>There are two cases:</p>\n\n<ul>\n<li>if $\\gamma =0$, then $\\alpha = f &gt;0$, and thus $r_1 = r_2$ follows from $(3)$. Then $(1)-(2)$ become the alternative $(a_i,b_i) =0$ or $a_i^2+ b_i^2 = \\alpha$. Since $a$ and $b$ are not proportional, the latter cas happens at least twice, and thus $2 = ||a||_2^2 + ||b||_2^2 \\geq 2 \\alpha$. Thus $f = \\alpha \\leq 1$, and in particular $f \\leq 1 + c^2$.</li>\n<li>if $\\gamma \\neq 0$ then there are at most $2$ projective solutions to $(3)$. Since the system $(1)-(2)$ is inhomogeneous, it has at most $2$ non-zero solutions up to sign. Thus there are non zero vectors $(a_1'',b_1'')$ and $(a_2'',b_2'')$ such that for eachi $i$ one has either $(a_i,b_i) = \\pm (a_j'',b_j'')$ for some $j=1,2$, or $(a_i,b_i) = 0$. Let $[|1,n|] = I_0 \\sqcup I_1 \\sqcup I_2$ such that $(a_i,b_i) = 0$ for $i$ in $I_0$, and $(a_i,b_i) = \\pm (a_j'',b_j'')$ if $i$ is in $I_j$ for some $j=1,2$. We then have the identities\n\\begin{align*}\n|I_1| (a_1'')^2 + |I_2| (a_2'')^2 &amp;= \\sum_i a_i^2 = 1 \\\\\n|I_1| a_1''b_1'' + |I_2| a_2'' b_2'' &amp;= \\sum_i a_i b_i = c \\\\\n|I_1| (b_1'')^2 + |I_2| (b_2'')^2 &amp;= \\sum_i b_i^2 = 1 \n\\end{align*}\nThus the vectors $a' = (|I_1|^{\\frac{1}{2}} a_1'',|I_2|^{\\frac{1}{2}}a_2'')$ and $b' = (|I_1|^{\\frac{1}{2}} b_1'',|I_2|^{\\frac{1}{2}}b_2'')$ in $\\mathbb{R}^2$ satisfy\n$$\n||a'||_2 =1, \\ \\ ||b'|| = 1, \\ \\ \\langle a',b' \\rangle = c.\n$$\nMoreover, one has\n\\begin{align*}\n\\sum_{i=1}^2 (a_i')^4 &amp;= |I_1|^2 (a_1'')^4 + |I_2|^2 (a_2'')^4 \\geq |I_1| (a_1'')^4 + |I_2| (a_2'')^4 = r_1^2 \\\\\n\\sum_{i=1}^2 (b_i')^4 &amp;= |I_1|^2 (b_1'')^4 + |I_2|^2 (b_2'')^4 \\geq |I_1| (b_1'')^4 + |I_2| (b_2'')^4 = r_2^2 \\\\\n\\sum_{i=1}^2 (a_i')^2(b_i')^2 &amp;= |I_1|^2 (a_1'')^2(b_1'')^2 + |I_2|^2 (a_2'')^2(b_2'')^2 \\geq |I_1| (a_1'')^2(b_1'')^2 + |I_2| (a_2'')^2 (b_2'')^2 = r_3 \\\\\n\\end{align*}\nConsequently, applying the case $n=2$ of the inequality to the vectors $a',b'$ yields\n$$\n1 + c^2 \\geq r_1 r_2 + r_3 = f(a,b).\n$$\nThus $f \\leq 1+c^2$ on all of $\\Sigma$.</li>\n</ul>\n" }, { "answer_id": 259934, "score": 5, "text": "<p>Just to clarify. As I remember, this is simply equivalent to the (partial case of) the question you cite. Since the cited question was solved, I would not call it a conjecture. But the question to find an independent proof makes sense. </p>\n\n<p>Well, let me elaborate. <a href=\"https://mathoverflow.net/q/234218\">Here</a> proving the partial case $N=2$ of your inequality I prove the following inequality: $$\\frac{\\|x\\|^2\\cdot \\|y\\|^2+(x,y)^2}{\\|Tx\\|^2 \\|Ty\\|^2}\\geqslant \\frac2{{\\rm tr}\\, T^4}$$\nfor any self-adjoint positive definite operator $T$ and any two vectors $x,y$ in $\\mathbb{R}^n$. If we denote $x=(a_1,\\dots,a_n)$, $y=(b_1,\\dots,b_n)$, $p=(a_1^2,\\dots,a_n^2)$, $q=(b_1^2,\\dots,b_n^2)$, $T^2=diag(s_1,\\dots,s_n)$, $s=(s_1,\\dots,s_n)$, we rewrite this as $$\\left(\\sum_{i=1}^n a_i^2\\right)\\left(\\sum_{i=1}^n b_i^2\\right)+\\left(\\sum_{i=1}^na_i b_i\\right)^2\\ge 2\\frac{(p,s)\\cdot (q,s)}{(s,s)},$$\nand maximizing over $s$ gives you $$\\sqrt{\\left(\\sum_{i=1}^n a_i^4\\right)\\left(\\sum_{i=1}^n b_i^4\\right)}+\\sum_{i=1}^na_i^2b_i^2$$ in RHS, this is the equality case in the lemma in the cited answer.</p>\n" }, { "answer_id": 259699, "score": 2, "text": "<p>Denote the power-sum (symmetric) polynomials by $p_k(a;n)=\\sum_{i=1}^na_i^k$; where $a=(a_1,\\dots, a_n)$ and also we need the Hadamard product $ab=(a_1b_1,\\dots,a_nb_n)$. Write $p_k(a)$ for $p_k(a;n)$ when there is no confusion. </p>\n\n<p><strong>Remark.</strong> Observe that $p_2(a)p_2(b)-p_2(ab)=\\sum_{i\\neq j}a_i^2b_j^2\\geq0$. We prove\n$$[p_2(a)p_2(b)+p_1(ab)^2-p_2(ab)]^2-p_2(a^2)p_2(b^2)\\geq0.\\tag1$$\nInduct on $n$. The case $n=2$ is noted as trivial. Assume true for $n$, we show for $n+1$; i.e.\n$$[(x^2+p_2(a))(y^2+p_2(b))+(xy+p_1(ab))^2-x^2y^2-p_2(ab)]^2-(x^4+p_2(a^2))(y^4+p_2(b^2))\\geq0.\\tag2$$\nAfter expansion, the LHS of (2) is a polynomial in even powers of $x$ and $y$. It suffices to study the following coefficients (the others follow by symmetry):</p>\n\n<p>$[x^0y^0]$: the is exactly the induction assumption, hence positive.</p>\n\n<p>$[x^2y^0]$: $2p_2(b)[p_1(ab)^2+p_2(a)p_2(b)-p_2(ab)]\\geq0$, by remark from above.</p>\n\n<p>$x^4y^0]$: $p_2(b)^2-p_2(b^2)\\geq0$, by remark.</p>\n\n<p>$[x^2y^2]$: $6p_1(ab)^2+4p_2(a)p_2(b)-2p_2(ab)\\geq0$, by remark.</p>\n\n<p>$[x^4y^2]$: $2p_2(b)\\geq0$.</p>\n\n<p>$[x^0y^4]$: $p_2(a)^2-p_2(a^2)\\geq0$, by remark.</p>\n\n<p>Therefore, the claim (1) is valid for all $n$.</p>\n\n<p>The intent of this method is to reveal that the inequality (1) is not as sharp as we may wish.</p>\n" }, { "answer_id": 259675, "score": 19, "text": "<p>The inequality is equivalent to\n$$\n\\left(\\sum_{i&gt;j} (a_ib_j+a_jb_i)^2+\\sum_{i} a^2_ib^2_i \\right)^2\\geq \\sum_{i} a^4_i \\sum_{i} b^4_i.\n$$\nThe left hand side is greater or equal to\n$$\n\\sum_i a_i^4b_i^4+\\sum_{i&gt;j} (a_ib_j+a_jb_i)^4+2(a_ib_j+a_jb_i)^2(a_i^2b_i^2+a_j^2b_j^2)+2a_i^2b_i^2a_j^2b_j^2\n$$\nAs\n$$\n(a_ib_j+a_jb_i)^4+2(a_ib_j+a_jb_i)^2(a_i^2b_i^2+a_j^2b_j^2)+2a_i^2b_i^2a_j^2b_j^2\\geq a_i^4b_j^4+a_j^4b_i^4.\n$$\nis equivalent to\n$$\n(a_ib_j+a_jb_i)^2(a_ib_i+a_jb_j)^2\\geq 0\n$$\nthe LHS is larger or equal to\n$$\n\\sum_i a_i^4b_i^4+\\sum_{i&gt;j} a_i^4b_j^4+a_j^4b_i^4=\\sum_{i} a^4_i \\sum_{i} b^4_i.\n$$</p>\n" }, { "answer_id": 259684, "score": 7, "text": "<p>I have an other solution, which not as slick Markus Sprecher's one, but I think the method, which is quite general, is interesting in itself.</p>\n\n<p>Fix $c \\in \\left]-1,1\\right[$ and $n \\geq 2$, and consider the compact manifold\n$$\n\\Sigma = \\{ (a,b) \\in S^{n-1} \\times S^{n-1} \\ | \\ \\langle a , b \\rangle = c \\}.\n$$\nIt has dimension $2n - 3$, and the tangent space at $(a,b)$ is the orthogonal in $\\mathbb{R}^{2n}$ to the 3-space spanned by $(a,0),(0,b),(b,a)$.</p>\n\n<p>Consider the functions\n$$\nr_1 = \\sqrt{\\sum_i a_i^4}, \\ \\ r_2 = \\sqrt{\\sum_i b_i^4}, \\ \\ r_3 = \\sum_{i} a_i^2 b_i^2\n$$\nas continuous functions on $\\Sigma$. Let $(a,b)$ be a point of $\\Sigma$ where $f = r_1 r_2 + r_3$ is maximal. We would like to show that $f(a,b) \\leq 1 + c^2$.</p>\n\n<p>The gradient of $f$ at $(a,b)$ must be of the form $\\alpha(a,0) + \\beta(0,b) + \\gamma(b,a)$ for some real numbers $\\alpha,\\beta,\\gamma$. This gives the following $2n$ equations:\n$$\n(1) \\ \\ a_i b_i^2 + a_i^3 \\frac{r_2}{r_1} = \\alpha a_i + \\gamma b_i \\\\\n(2) \\ \\ b_i a_i^2 + b_i^3 \\frac{r_1}{r_2} = \\beta b_i + \\gamma a_i \\\\\n$$\nMultiplying $(1)$ by $a_i$ and summing ober $i$, we get $f = r_3 +r_1r_2 = \\alpha + c \\gamma$. \nMultiplying $(2)$ by $b_i$ and summing ober $i$, we get $f = r_3 +r_1r_2 = \\beta + c \\gamma$. In particular $\\alpha = \\beta$.</p>\n\n<p>Mutliply equations $(1)$ and $(2)$ by $r_1 b_i$ and $r_2 a_i$ respectively. This yields\n$$\n(1)' \\ \\ r_1 a_i b_i^3 + r_2 a_i^3 b_i = r_1 \\alpha a_i b_i + r_1 \\gamma b_i^2 \\\\\n(2)' \\ \\ r_2 a_i^3 b_i + r_1 a_i b_i^3 = r_2 \\alpha a_ib_i + r_2 \\gamma a_i^2 \\\\\n$$\nThe two LHSs are equal, hence so are the RHSs. Thus $(a_i,b_i)$ satisifies the quadratic equation\n$$\n(3) r_1 \\alpha a_i b_i + r_1 \\gamma b_i^2 = r_2 \\alpha a_ib_i + r_2 \\gamma a_i^2.\n$$</p>\n\n<p>There are two cases:</p>\n\n<ul>\n<li>if $\\gamma =0$, then $\\alpha = f &gt;0$, and thus $r_1 = r_2$ follows from $(3)$. Then $(1)-(2)$ become the alternative $(a_i,b_i) =0$ or $a_i^2+ b_i^2 = \\alpha$. Since $a$ and $b$ are not proportional, the latter cas happens at least twice, and thus $2 = ||a||_2^2 + ||b||_2^2 \\geq 2 \\alpha$. Thus $f = \\alpha \\leq 1$, and in particular $f \\leq 1 + c^2$.</li>\n<li>if $\\gamma \\neq 0$ then there are at most $2$ projective solutions to $(3)$. Since the system $(1)-(2)$ is inhomogeneous, it has at most $2$ non-zero solutions up to sign. Thus there are non zero vectors $(a_1'',b_1'')$ and $(a_2'',b_2'')$ such that for eachi $i$ one has either $(a_i,b_i) = \\pm (a_j'',b_j'')$ for some $j=1,2$, or $(a_i,b_i) = 0$. Let $[|1,n|] = I_0 \\sqcup I_1 \\sqcup I_2$ such that $(a_i,b_i) = 0$ for $i$ in $I_0$, and $(a_i,b_i) = \\pm (a_j'',b_j'')$ if $i$ is in $I_j$ for some $j=1,2$. We then have the identities\n\\begin{align*}\n|I_1| (a_1'')^2 + |I_2| (a_2'')^2 &amp;= \\sum_i a_i^2 = 1 \\\\\n|I_1| a_1''b_1'' + |I_2| a_2'' b_2'' &amp;= \\sum_i a_i b_i = c \\\\\n|I_1| (b_1'')^2 + |I_2| (b_2'')^2 &amp;= \\sum_i b_i^2 = 1 \n\\end{align*}\nThus the vectors $a' = (|I_1|^{\\frac{1}{2}} a_1'',|I_2|^{\\frac{1}{2}}a_2'')$ and $b' = (|I_1|^{\\frac{1}{2}} b_1'',|I_2|^{\\frac{1}{2}}b_2'')$ in $\\mathbb{R}^2$ satisfy\n$$\n||a'||_2 =1, \\ \\ ||b'|| = 1, \\ \\ \\langle a',b' \\rangle = c.\n$$\nMoreover, one has\n\\begin{align*}\n\\sum_{i=1}^2 (a_i')^4 &amp;= |I_1|^2 (a_1'')^4 + |I_2|^2 (a_2'')^4 \\geq |I_1| (a_1'')^4 + |I_2| (a_2'')^4 = r_1^2 \\\\\n\\sum_{i=1}^2 (b_i')^4 &amp;= |I_1|^2 (b_1'')^4 + |I_2|^2 (b_2'')^4 \\geq |I_1| (b_1'')^4 + |I_2| (b_2'')^4 = r_2^2 \\\\\n\\sum_{i=1}^2 (a_i')^2(b_i')^2 &amp;= |I_1|^2 (a_1'')^2(b_1'')^2 + |I_2|^2 (a_2'')^2(b_2'')^2 \\geq |I_1| (a_1'')^2(b_1'')^2 + |I_2| (a_2'')^2 (b_2'')^2 = r_3 \\\\\n\\end{align*}\nConsequently, applying the case $n=2$ of the inequality to the vectors $a',b'$ yields\n$$\n1 + c^2 \\geq r_1 r_2 + r_3 = f(a,b).\n$$\nThus $f \\leq 1+c^2$ on all of $\\Sigma$.</li>\n</ul>\n" }, { "answer_id": 259934, "score": 5, "text": "<p>Just to clarify. As I remember, this is simply equivalent to the (partial case of) the question you cite. Since the cited question was solved, I would not call it a conjecture. But the question to find an independent proof makes sense. </p>\n\n<p>Well, let me elaborate. <a href=\"https://mathoverflow.net/q/234218\">Here</a> proving the partial case $N=2$ of your inequality I prove the following inequality: $$\\frac{\\|x\\|^2\\cdot \\|y\\|^2+(x,y)^2}{\\|Tx\\|^2 \\|Ty\\|^2}\\geqslant \\frac2{{\\rm tr}\\, T^4}$$\nfor any self-adjoint positive definite operator $T$ and any two vectors $x,y$ in $\\mathbb{R}^n$. If we denote $x=(a_1,\\dots,a_n)$, $y=(b_1,\\dots,b_n)$, $p=(a_1^2,\\dots,a_n^2)$, $q=(b_1^2,\\dots,b_n^2)$, $T^2=diag(s_1,\\dots,s_n)$, $s=(s_1,\\dots,s_n)$, we rewrite this as $$\\left(\\sum_{i=1}^n a_i^2\\right)\\left(\\sum_{i=1}^n b_i^2\\right)+\\left(\\sum_{i=1}^na_i b_i\\right)^2\\ge 2\\frac{(p,s)\\cdot (q,s)}{(s,s)},$$\nand maximizing over $s$ gives you $$\\sqrt{\\left(\\sum_{i=1}^n a_i^4\\right)\\left(\\sum_{i=1}^n b_i^4\\right)}+\\sum_{i=1}^na_i^2b_i^2$$ in RHS, this is the equality case in the lemma in the cited answer.</p>\n" }, { "answer_id": 259699, "score": 2, "text": "<p>Denote the power-sum (symmetric) polynomials by $p_k(a;n)=\\sum_{i=1}^na_i^k$; where $a=(a_1,\\dots, a_n)$ and also we need the Hadamard product $ab=(a_1b_1,\\dots,a_nb_n)$. Write $p_k(a)$ for $p_k(a;n)$ when there is no confusion. </p>\n\n<p><strong>Remark.</strong> Observe that $p_2(a)p_2(b)-p_2(ab)=\\sum_{i\\neq j}a_i^2b_j^2\\geq0$. We prove\n$$[p_2(a)p_2(b)+p_1(ab)^2-p_2(ab)]^2-p_2(a^2)p_2(b^2)\\geq0.\\tag1$$\nInduct on $n$. The case $n=2$ is noted as trivial. Assume true for $n$, we show for $n+1$; i.e.\n$$[(x^2+p_2(a))(y^2+p_2(b))+(xy+p_1(ab))^2-x^2y^2-p_2(ab)]^2-(x^4+p_2(a^2))(y^4+p_2(b^2))\\geq0.\\tag2$$\nAfter expansion, the LHS of (2) is a polynomial in even powers of $x$ and $y$. It suffices to study the following coefficients (the others follow by symmetry):</p>\n\n<p>$[x^0y^0]$: the is exactly the induction assumption, hence positive.</p>\n\n<p>$[x^2y^0]$: $2p_2(b)[p_1(ab)^2+p_2(a)p_2(b)-p_2(ab)]\\geq0$, by remark from above.</p>\n\n<p>$x^4y^0]$: $p_2(b)^2-p_2(b^2)\\geq0$, by remark.</p>\n\n<p>$[x^2y^2]$: $6p_1(ab)^2+4p_2(a)p_2(b)-2p_2(ab)\\geq0$, by remark.</p>\n\n<p>$[x^4y^2]$: $2p_2(b)\\geq0$.</p>\n\n<p>$[x^0y^4]$: $p_2(a)^2-p_2(a^2)\\geq0$, by remark.</p>\n\n<p>Therefore, the claim (1) is valid for all $n$.</p>\n\n<p>The intent of this method is to reveal that the inequality (1) is not as sharp as we may wish.</p>\n" } ]
Question: A seemingly simple inequality Answer 1: <p>The inequality is equivalent to $$ \left(\sum_{i&gt;j} (a_ib_j+a_jb_i)^2+\sum_{i} a^2_ib^2_i \right)^2\geq \sum_{i} a^4_i \sum_{i} b^4_i. $$ The left hand side is greater or equal to $$ \sum_i a_i^4b_i^4+\sum_{i&gt;j} (a_ib_j+a_jb_i)^4+2(a_ib_j+a_jb_i)^2(a_i^2b_i^2+a_j^2b_j^2)+2a_i^2b_i^2a_j^2b_j^2 $$ As $$ (a_ib_j+a_jb_i)^4+2(a_ib_j+a_jb_i)^2(a_i^2b_i^2+a_j^2b_j^2)+2a_i^2b_i^2a_j^2b_j^2\geq a_i^4b_j^4+a_j^4b_i^4. $$ is equivalent to $$ (a_ib_j+a_jb_i)^2(a_ib_i+a_jb_j)^2\geq 0 $$ the LHS is larger or equal to $$ \sum_i a_i^4b_i^4+\sum_{i&gt;j} a_i^4b_j^4+a_j^4b_i^4=\sum_{i} a^4_i \sum_{i} b^4_i. $$</p> Answer 2: <p>I have an other solution, which not as slick Markus Sprecher's one, but I think the method, which is quite general, is interesting in itself.</p> <p>Fix $c \in \left]-1,1\right[$ and $n \geq 2$, and consider the compact manifold $$ \Sigma = \{ (a,b) \in S^{n-1} \times S^{n-1} \ | \ \langle a , b \rangle = c \}. $$ It has dimension $2n - 3$, and the tangent space at $(a,b)$ is the orthogonal in $\mathbb{R}^{2n}$ to the 3-space spanned by $(a,0),(0,b),(b,a)$.</p> <p>Consider the functions $$ r_1 = \sqrt{\sum_i a_i^4}, \ \ r_2 = \sqrt{\sum_i b_i^4}, \ \ r_3 = \sum_{i} a_i^2 b_i^2 $$ as continuous functions on $\Sigma$. Let $(a,b)$ be a point of $\Sigma$ where $f = r_1 r_2 + r_3$ is maximal. We would like to show that $f(a,b) \leq 1 + c^2$.</p> <p>The gradient of $f$ at $(a,b)$ must be of the form $\alpha(a,0) + \beta(0,b) + \gamma(b,a)$ for some real numbers $\alpha,\beta,\gamma$. This gives the following $2n$ equations: $$ (1) \ \ a_i b_i^2 + a_i^3 \frac{r_2}{r_1} = \alpha a_i + \gamma b_i \\ (2) \ \ b_i a_i^2 + b_i^3 \frac{r_1}{r_2} = \beta b_i + \gamma a_i \\ $$ Multiplying $(1)$ by $a_i$ and summing ober $i$, we get $f = r_3 +r_1r_2 = \alpha + c \gamma$. Multiplying $(2)$ by $b_i$ and summing ober $i$, we get $f = r_3 +r_1r_2 = \beta + c \gamma$. In particular $\alpha = \beta$.</p> <p>Mutliply equations $(1)$ and $(2)$ by $r_1 b_i$ and $r_2 a_i$ respectively. This yields $$ (1)' \ \ r_1 a_i b_i^3 + r_2 a_i^3 b_i = r_1 \alpha a_i b_i + r_1 \gamma b_i^2 \\ (2)' \ \ r_2 a_i^3 b_i + r_1 a_i b_i^3 = r_2 \alpha a_ib_i + r_2 \gamma a_i^2 \\ $$ The two LHSs are equal, hence so are the RHSs. Thus $(a_i,b_i)$ satisifies the quadratic equation $$ (3) r_1 \alpha a_i b_i + r_1 \gamma b_i^2 = r_2 \alpha a_ib_i + r_2 \gamma a_i^2. $$</p> <p>There are two cases:</p> <ul> <li>if $\gamma =0$, then $\alpha = f &gt;0$, and thus $r_1 = r_2$ follows from $(3)$. Then $(1)-(2)$ become the alternative $(a_i,b_i) =0$ or $a_i^2+ b_i^2 = \alpha$. Since $a$ and $b$ are not proportional, the latter cas happens at least twice, and thus $2 = ||a||_2^2 + ||b||_2^2 \geq 2 \alpha$. Thus $f = \alpha \leq 1$, and in particular $f \leq 1 + c^2$.</li> <li>if $\gamma \neq 0$ then there are at most $2$ projective solutions to $(3)$. Since the system $(1)-(2)$ is inhomogeneous, it has at most $2$ non-zero solutions up to sign. Thus there are non zero vectors $(a_1'',b_1'')$ and $(a_2'',b_2'')$ such that for eachi $i$ one has either $(a_i,b_i) = \pm (a_j'',b_j'')$ for some $j=1,2$, or $(a_i,b_i) = 0$. Let $[|1,n|] = I_0 \sqcup I_1 \sqcup I_2$ such that $(a_i,b_i) = 0$ for $i$ in $I_0$, and $(a_i,b_i) = \pm (a_j'',b_j'')$ if $i$ is in $I_j$ for some $j=1,2$. We then have the identities \begin{align*} |I_1| (a_1'')^2 + |I_2| (a_2'')^2 &amp;= \sum_i a_i^2 = 1 \\ |I_1| a_1''b_1'' + |I_2| a_2'' b_2'' &amp;= \sum_i a_i b_i = c \\ |I_1| (b_1'')^2 + |I_2| (b_2'')^2 &amp;= \sum_i b_i^2 = 1 \end{align*} Thus the vectors $a' = (|I_1|^{\frac{1}{2}} a_1'',|I_2|^{\frac{1}{2}}a_2'')$ and $b' = (|I_1|^{\frac{1}{2}} b_1'',|I_2|^{\frac{1}{2}}b_2'')$ in $\mathbb{R}^2$ satisfy $$ ||a'||_2 =1, \ \ ||b'|| = 1, \ \ \langle a',b' \rangle = c. $$ Moreover, one has \begin{align*} \sum_{i=1}^2 (a_i')^4 &amp;= |I_1|^2 (a_1'')^4 + |I_2|^2 (a_2'')^4 \geq |I_1| (a_1'')^4 + |I_2| (a_2'')^4 = r_1^2 \\ \sum_{i=1}^2 (b_i')^4 &amp;= |I_1|^2 (b_1'')^4 + |I_2|^2 (b_2'')^4 \geq |I_1| (b_1'')^4 + |I_2| (b_2'')^4 = r_2^2 \\ \sum_{i=1}^2 (a_i')^2(b_i')^2 &amp;= |I_1|^2 (a_1'')^2(b_1'')^2 + |I_2|^2 (a_2'')^2(b_2'')^2 \geq |I_1| (a_1'')^2(b_1'')^2 + |I_2| (a_2'')^2 (b_2'')^2 = r_3 \\ \end{align*} Consequently, applying the case $n=2$ of the inequality to the vectors $a',b'$ yields $$ 1 + c^2 \geq r_1 r_2 + r_3 = f(a,b). $$ Thus $f \leq 1+c^2$ on all of $\Sigma$.</li> </ul> Answer 3: <p>Just to clarify. As I remember, this is simply equivalent to the (partial case of) the question you cite. Since the cited question was solved, I would not call it a conjecture. But the question to find an independent proof makes sense. </p> <p>Well, let me elaborate. <a href="https://mathoverflow.net/q/234218">Here</a> proving the partial case $N=2$ of your inequality I prove the following inequality: $$\frac{\|x\|^2\cdot \|y\|^2+(x,y)^2}{\|Tx\|^2 \|Ty\|^2}\geqslant \frac2{{\rm tr}\, T^4}$$ for any self-adjoint positive definite operator $T$ and any two vectors $x,y$ in $\mathbb{R}^n$. If we denote $x=(a_1,\dots,a_n)$, $y=(b_1,\dots,b_n)$, $p=(a_1^2,\dots,a_n^2)$, $q=(b_1^2,\dots,b_n^2)$, $T^2=diag(s_1,\dots,s_n)$, $s=(s_1,\dots,s_n)$, we rewrite this as $$\left(\sum_{i=1}^n a_i^2\right)\left(\sum_{i=1}^n b_i^2\right)+\left(\sum_{i=1}^na_i b_i\right)^2\ge 2\frac{(p,s)\cdot (q,s)}{(s,s)},$$ and maximizing over $s$ gives you $$\sqrt{\left(\sum_{i=1}^n a_i^4\right)\left(\sum_{i=1}^n b_i^4\right)}+\sum_{i=1}^na_i^2b_i^2$$ in RHS, this is the equality case in the lemma in the cited answer.</p> Answer 4: <p>Denote the power-sum (symmetric) polynomials by $p_k(a;n)=\sum_{i=1}^na_i^k$; where $a=(a_1,\dots, a_n)$ and also we need the Hadamard product $ab=(a_1b_1,\dots,a_nb_n)$. Write $p_k(a)$ for $p_k(a;n)$ when there is no confusion. </p> <p><strong>Remark.</strong> Observe that $p_2(a)p_2(b)-p_2(ab)=\sum_{i\neq j}a_i^2b_j^2\geq0$. We prove $$[p_2(a)p_2(b)+p_1(ab)^2-p_2(ab)]^2-p_2(a^2)p_2(b^2)\geq0.\tag1$$ Induct on $n$. The case $n=2$ is noted as trivial. Assume true for $n$, we show for $n+1$; i.e. $$[(x^2+p_2(a))(y^2+p_2(b))+(xy+p_1(ab))^2-x^2y^2-p_2(ab)]^2-(x^4+p_2(a^2))(y^4+p_2(b^2))\geq0.\tag2$$ After expansion, the LHS of (2) is a polynomial in even powers of $x$ and $y$. It suffices to study the following coefficients (the others follow by symmetry):</p> <p>$[x^0y^0]$: the is exactly the induction assumption, hence positive.</p> <p>$[x^2y^0]$: $2p_2(b)[p_1(ab)^2+p_2(a)p_2(b)-p_2(ab)]\geq0$, by remark from above.</p> <p>$x^4y^0]$: $p_2(b)^2-p_2(b^2)\geq0$, by remark.</p> <p>$[x^2y^2]$: $6p_1(ab)^2+4p_2(a)p_2(b)-2p_2(ab)\geq0$, by remark.</p> <p>$[x^4y^2]$: $2p_2(b)\geq0$.</p> <p>$[x^0y^4]$: $p_2(a)^2-p_2(a^2)\geq0$, by remark.</p> <p>Therefore, the claim (1) is valid for all $n$.</p> <p>The intent of this method is to reveal that the inequality (1) is not as sharp as we may wish.</p> Answer 5: <p>The inequality is equivalent to $$ \left(\sum_{i&gt;j} (a_ib_j+a_jb_i)^2+\sum_{i} a^2_ib^2_i \right)^2\geq \sum_{i} a^4_i \sum_{i} b^4_i. $$ The left hand side is greater or equal to $$ \sum_i a_i^4b_i^4+\sum_{i&gt;j} (a_ib_j+a_jb_i)^4+2(a_ib_j+a_jb_i)^2(a_i^2b_i^2+a_j^2b_j^2)+2a_i^2b_i^2a_j^2b_j^2 $$ As $$ (a_ib_j+a_jb_i)^4+2(a_ib_j+a_jb_i)^2(a_i^2b_i^2+a_j^2b_j^2)+2a_i^2b_i^2a_j^2b_j^2\geq a_i^4b_j^4+a_j^4b_i^4. $$ is equivalent to $$ (a_ib_j+a_jb_i)^2(a_ib_i+a_jb_j)^2\geq 0 $$ the LHS is larger or equal to $$ \sum_i a_i^4b_i^4+\sum_{i&gt;j} a_i^4b_j^4+a_j^4b_i^4=\sum_{i} a^4_i \sum_{i} b^4_i. $$</p> Answer 6: <p>I have an other solution, which not as slick Markus Sprecher's one, but I think the method, which is quite general, is interesting in itself.</p> <p>Fix $c \in \left]-1,1\right[$ and $n \geq 2$, and consider the compact manifold $$ \Sigma = \{ (a,b) \in S^{n-1} \times S^{n-1} \ | \ \langle a , b \rangle = c \}. $$ It has dimension $2n - 3$, and the tangent space at $(a,b)$ is the orthogonal in $\mathbb{R}^{2n}$ to the 3-space spanned by $(a,0),(0,b),(b,a)$.</p> <p>Consider the functions $$ r_1 = \sqrt{\sum_i a_i^4}, \ \ r_2 = \sqrt{\sum_i b_i^4}, \ \ r_3 = \sum_{i} a_i^2 b_i^2 $$ as continuous functions on $\Sigma$. Let $(a,b)$ be a point of $\Sigma$ where $f = r_1 r_2 + r_3$ is maximal. We would like to show that $f(a,b) \leq 1 + c^2$.</p> <p>The gradient of $f$ at $(a,b)$ must be of the form $\alpha(a,0) + \beta(0,b) + \gamma(b,a)$ for some real numbers $\alpha,\beta,\gamma$. This gives the following $2n$ equations: $$ (1) \ \ a_i b_i^2 + a_i^3 \frac{r_2}{r_1} = \alpha a_i + \gamma b_i \\ (2) \ \ b_i a_i^2 + b_i^3 \frac{r_1}{r_2} = \beta b_i + \gamma a_i \\ $$ Multiplying $(1)$ by $a_i$ and summing ober $i$, we get $f = r_3 +r_1r_2 = \alpha + c \gamma$. Multiplying $(2)$ by $b_i$ and summing ober $i$, we get $f = r_3 +r_1r_2 = \beta + c \gamma$. In particular $\alpha = \beta$.</p> <p>Mutliply equations $(1)$ and $(2)$ by $r_1 b_i$ and $r_2 a_i$ respectively. This yields $$ (1)' \ \ r_1 a_i b_i^3 + r_2 a_i^3 b_i = r_1 \alpha a_i b_i + r_1 \gamma b_i^2 \\ (2)' \ \ r_2 a_i^3 b_i + r_1 a_i b_i^3 = r_2 \alpha a_ib_i + r_2 \gamma a_i^2 \\ $$ The two LHSs are equal, hence so are the RHSs. Thus $(a_i,b_i)$ satisifies the quadratic equation $$ (3) r_1 \alpha a_i b_i + r_1 \gamma b_i^2 = r_2 \alpha a_ib_i + r_2 \gamma a_i^2. $$</p> <p>There are two cases:</p> <ul> <li>if $\gamma =0$, then $\alpha = f &gt;0$, and thus $r_1 = r_2$ follows from $(3)$. Then $(1)-(2)$ become the alternative $(a_i,b_i) =0$ or $a_i^2+ b_i^2 = \alpha$. Since $a$ and $b$ are not proportional, the latter cas happens at least twice, and thus $2 = ||a||_2^2 + ||b||_2^2 \geq 2 \alpha$. Thus $f = \alpha \leq 1$, and in particular $f \leq 1 + c^2$.</li> <li>if $\gamma \neq 0$ then there are at most $2$ projective solutions to $(3)$. Since the system $(1)-(2)$ is inhomogeneous, it has at most $2$ non-zero solutions up to sign. Thus there are non zero vectors $(a_1'',b_1'')$ and $(a_2'',b_2'')$ such that for eachi $i$ one has either $(a_i,b_i) = \pm (a_j'',b_j'')$ for some $j=1,2$, or $(a_i,b_i) = 0$. Let $[|1,n|] = I_0 \sqcup I_1 \sqcup I_2$ such that $(a_i,b_i) = 0$ for $i$ in $I_0$, and $(a_i,b_i) = \pm (a_j'',b_j'')$ if $i$ is in $I_j$ for some $j=1,2$. We then have the identities \begin{align*} |I_1| (a_1'')^2 + |I_2| (a_2'')^2 &amp;= \sum_i a_i^2 = 1 \\ |I_1| a_1''b_1'' + |I_2| a_2'' b_2'' &amp;= \sum_i a_i b_i = c \\ |I_1| (b_1'')^2 + |I_2| (b_2'')^2 &amp;= \sum_i b_i^2 = 1 \end{align*} Thus the vectors $a' = (|I_1|^{\frac{1}{2}} a_1'',|I_2|^{\frac{1}{2}}a_2'')$ and $b' = (|I_1|^{\frac{1}{2}} b_1'',|I_2|^{\frac{1}{2}}b_2'')$ in $\mathbb{R}^2$ satisfy $$ ||a'||_2 =1, \ \ ||b'|| = 1, \ \ \langle a',b' \rangle = c. $$ Moreover, one has \begin{align*} \sum_{i=1}^2 (a_i')^4 &amp;= |I_1|^2 (a_1'')^4 + |I_2|^2 (a_2'')^4 \geq |I_1| (a_1'')^4 + |I_2| (a_2'')^4 = r_1^2 \\ \sum_{i=1}^2 (b_i')^4 &amp;= |I_1|^2 (b_1'')^4 + |I_2|^2 (b_2'')^4 \geq |I_1| (b_1'')^4 + |I_2| (b_2'')^4 = r_2^2 \\ \sum_{i=1}^2 (a_i')^2(b_i')^2 &amp;= |I_1|^2 (a_1'')^2(b_1'')^2 + |I_2|^2 (a_2'')^2(b_2'')^2 \geq |I_1| (a_1'')^2(b_1'')^2 + |I_2| (a_2'')^2 (b_2'')^2 = r_3 \\ \end{align*} Consequently, applying the case $n=2$ of the inequality to the vectors $a',b'$ yields $$ 1 + c^2 \geq r_1 r_2 + r_3 = f(a,b). $$ Thus $f \leq 1+c^2$ on all of $\Sigma$.</li> </ul> Answer 7: <p>Just to clarify. As I remember, this is simply equivalent to the (partial case of) the question you cite. Since the cited question was solved, I would not call it a conjecture. But the question to find an independent proof makes sense. </p> <p>Well, let me elaborate. <a href="https://mathoverflow.net/q/234218">Here</a> proving the partial case $N=2$ of your inequality I prove the following inequality: $$\frac{\|x\|^2\cdot \|y\|^2+(x,y)^2}{\|Tx\|^2 \|Ty\|^2}\geqslant \frac2{{\rm tr}\, T^4}$$ for any self-adjoint positive definite operator $T$ and any two vectors $x,y$ in $\mathbb{R}^n$. If we denote $x=(a_1,\dots,a_n)$, $y=(b_1,\dots,b_n)$, $p=(a_1^2,\dots,a_n^2)$, $q=(b_1^2,\dots,b_n^2)$, $T^2=diag(s_1,\dots,s_n)$, $s=(s_1,\dots,s_n)$, we rewrite this as $$\left(\sum_{i=1}^n a_i^2\right)\left(\sum_{i=1}^n b_i^2\right)+\left(\sum_{i=1}^na_i b_i\right)^2\ge 2\frac{(p,s)\cdot (q,s)}{(s,s)},$$ and maximizing over $s$ gives you $$\sqrt{\left(\sum_{i=1}^n a_i^4\right)\left(\sum_{i=1}^n b_i^4\right)}+\sum_{i=1}^na_i^2b_i^2$$ in RHS, this is the equality case in the lemma in the cited answer.</p> Answer 8: <p>Denote the power-sum (symmetric) polynomials by $p_k(a;n)=\sum_{i=1}^na_i^k$; where $a=(a_1,\dots, a_n)$ and also we need the Hadamard product $ab=(a_1b_1,\dots,a_nb_n)$. Write $p_k(a)$ for $p_k(a;n)$ when there is no confusion. </p> <p><strong>Remark.</strong> Observe that $p_2(a)p_2(b)-p_2(ab)=\sum_{i\neq j}a_i^2b_j^2\geq0$. We prove $$[p_2(a)p_2(b)+p_1(ab)^2-p_2(ab)]^2-p_2(a^2)p_2(b^2)\geq0.\tag1$$ Induct on $n$. The case $n=2$ is noted as trivial. Assume true for $n$, we show for $n+1$; i.e. $$[(x^2+p_2(a))(y^2+p_2(b))+(xy+p_1(ab))^2-x^2y^2-p_2(ab)]^2-(x^4+p_2(a^2))(y^4+p_2(b^2))\geq0.\tag2$$ After expansion, the LHS of (2) is a polynomial in even powers of $x$ and $y$. It suffices to study the following coefficients (the others follow by symmetry):</p> <p>$[x^0y^0]$: the is exactly the induction assumption, hence positive.</p> <p>$[x^2y^0]$: $2p_2(b)[p_1(ab)^2+p_2(a)p_2(b)-p_2(ab)]\geq0$, by remark from above.</p> <p>$x^4y^0]$: $p_2(b)^2-p_2(b^2)\geq0$, by remark.</p> <p>$[x^2y^2]$: $6p_1(ab)^2+4p_2(a)p_2(b)-2p_2(ab)\geq0$, by remark.</p> <p>$[x^4y^2]$: $2p_2(b)\geq0$.</p> <p>$[x^0y^4]$: $p_2(a)^2-p_2(a^2)\geq0$, by remark.</p> <p>Therefore, the claim (1) is valid for all $n$.</p> <p>The intent of this method is to reveal that the inequality (1) is not as sharp as we may wish.</p> Answer 9: <p>The inequality is equivalent to $$ \left(\sum_{i&gt;j} (a_ib_j+a_jb_i)^2+\sum_{i} a^2_ib^2_i \right)^2\geq \sum_{i} a^4_i \sum_{i} b^4_i. $$ The left hand side is greater or equal to $$ \sum_i a_i^4b_i^4+\sum_{i&gt;j} (a_ib_j+a_jb_i)^4+2(a_ib_j+a_jb_i)^2(a_i^2b_i^2+a_j^2b_j^2)+2a_i^2b_i^2a_j^2b_j^2 $$ As $$ (a_ib_j+a_jb_i)^4+2(a_ib_j+a_jb_i)^2(a_i^2b_i^2+a_j^2b_j^2)+2a_i^2b_i^2a_j^2b_j^2\geq a_i^4b_j^4+a_j^4b_i^4. $$ is equivalent to $$ (a_ib_j+a_jb_i)^2(a_ib_i+a_jb_j)^2\geq 0 $$ the LHS is larger or equal to $$ \sum_i a_i^4b_i^4+\sum_{i&gt;j} a_i^4b_j^4+a_j^4b_i^4=\sum_{i} a^4_i \sum_{i} b^4_i. $$</p> Answer 10: <p>I have an other solution, which not as slick Markus Sprecher's one, but I think the method, which is quite general, is interesting in itself.</p> <p>Fix $c \in \left]-1,1\right[$ and $n \geq 2$, and consider the compact manifold $$ \Sigma = \{ (a,b) \in S^{n-1} \times S^{n-1} \ | \ \langle a , b \rangle = c \}. $$ It has dimension $2n - 3$, and the tangent space at $(a,b)$ is the orthogonal in $\mathbb{R}^{2n}$ to the 3-space spanned by $(a,0),(0,b),(b,a)$.</p> <p>Consider the functions $$ r_1 = \sqrt{\sum_i a_i^4}, \ \ r_2 = \sqrt{\sum_i b_i^4}, \ \ r_3 = \sum_{i} a_i^2 b_i^2 $$ as continuous functions on $\Sigma$. Let $(a,b)$ be a point of $\Sigma$ where $f = r_1 r_2 + r_3$ is maximal. We would like to show that $f(a,b) \leq 1 + c^2$.</p> <p>The gradient of $f$ at $(a,b)$ must be of the form $\alpha(a,0) + \beta(0,b) + \gamma(b,a)$ for some real numbers $\alpha,\beta,\gamma$. This gives the following $2n$ equations: $$ (1) \ \ a_i b_i^2 + a_i^3 \frac{r_2}{r_1} = \alpha a_i + \gamma b_i \\ (2) \ \ b_i a_i^2 + b_i^3 \frac{r_1}{r_2} = \beta b_i + \gamma a_i \\ $$ Multiplying $(1)$ by $a_i$ and summing ober $i$, we get $f = r_3 +r_1r_2 = \alpha + c \gamma$. Multiplying $(2)$ by $b_i$ and summing ober $i$, we get $f = r_3 +r_1r_2 = \beta + c \gamma$. In particular $\alpha = \beta$.</p> <p>Mutliply equations $(1)$ and $(2)$ by $r_1 b_i$ and $r_2 a_i$ respectively. This yields $$ (1)' \ \ r_1 a_i b_i^3 + r_2 a_i^3 b_i = r_1 \alpha a_i b_i + r_1 \gamma b_i^2 \\ (2)' \ \ r_2 a_i^3 b_i + r_1 a_i b_i^3 = r_2 \alpha a_ib_i + r_2 \gamma a_i^2 \\ $$ The two LHSs are equal, hence so are the RHSs. Thus $(a_i,b_i)$ satisifies the quadratic equation $$ (3) r_1 \alpha a_i b_i + r_1 \gamma b_i^2 = r_2 \alpha a_ib_i + r_2 \gamma a_i^2. $$</p> <p>There are two cases:</p> <ul> <li>if $\gamma =0$, then $\alpha = f &gt;0$, and thus $r_1 = r_2$ follows from $(3)$. Then $(1)-(2)$ become the alternative $(a_i,b_i) =0$ or $a_i^2+ b_i^2 = \alpha$. Since $a$ and $b$ are not proportional, the latter cas happens at least twice, and thus $2 = ||a||_2^2 + ||b||_2^2 \geq 2 \alpha$. Thus $f = \alpha \leq 1$, and in particular $f \leq 1 + c^2$.</li> <li>if $\gamma \neq 0$ then there are at most $2$ projective solutions to $(3)$. Since the system $(1)-(2)$ is inhomogeneous, it has at most $2$ non-zero solutions up to sign. Thus there are non zero vectors $(a_1'',b_1'')$ and $(a_2'',b_2'')$ such that for eachi $i$ one has either $(a_i,b_i) = \pm (a_j'',b_j'')$ for some $j=1,2$, or $(a_i,b_i) = 0$. Let $[|1,n|] = I_0 \sqcup I_1 \sqcup I_2$ such that $(a_i,b_i) = 0$ for $i$ in $I_0$, and $(a_i,b_i) = \pm (a_j'',b_j'')$ if $i$ is in $I_j$ for some $j=1,2$. We then have the identities \begin{align*} |I_1| (a_1'')^2 + |I_2| (a_2'')^2 &amp;= \sum_i a_i^2 = 1 \\ |I_1| a_1''b_1'' + |I_2| a_2'' b_2'' &amp;= \sum_i a_i b_i = c \\ |I_1| (b_1'')^2 + |I_2| (b_2'')^2 &amp;= \sum_i b_i^2 = 1 \end{align*} Thus the vectors $a' = (|I_1|^{\frac{1}{2}} a_1'',|I_2|^{\frac{1}{2}}a_2'')$ and $b' = (|I_1|^{\frac{1}{2}} b_1'',|I_2|^{\frac{1}{2}}b_2'')$ in $\mathbb{R}^2$ satisfy $$ ||a'||_2 =1, \ \ ||b'|| = 1, \ \ \langle a',b' \rangle = c. $$ Moreover, one has \begin{align*} \sum_{i=1}^2 (a_i')^4 &amp;= |I_1|^2 (a_1'')^4 + |I_2|^2 (a_2'')^4 \geq |I_1| (a_1'')^4 + |I_2| (a_2'')^4 = r_1^2 \\ \sum_{i=1}^2 (b_i')^4 &amp;= |I_1|^2 (b_1'')^4 + |I_2|^2 (b_2'')^4 \geq |I_1| (b_1'')^4 + |I_2| (b_2'')^4 = r_2^2 \\ \sum_{i=1}^2 (a_i')^2(b_i')^2 &amp;= |I_1|^2 (a_1'')^2(b_1'')^2 + |I_2|^2 (a_2'')^2(b_2'')^2 \geq |I_1| (a_1'')^2(b_1'')^2 + |I_2| (a_2'')^2 (b_2'')^2 = r_3 \\ \end{align*} Consequently, applying the case $n=2$ of the inequality to the vectors $a',b'$ yields $$ 1 + c^2 \geq r_1 r_2 + r_3 = f(a,b). $$ Thus $f \leq 1+c^2$ on all of $\Sigma$.</li> </ul> Answer 11: <p>Just to clarify. As I remember, this is simply equivalent to the (partial case of) the question you cite. Since the cited question was solved, I would not call it a conjecture. But the question to find an independent proof makes sense. </p> <p>Well, let me elaborate. <a href="https://mathoverflow.net/q/234218">Here</a> proving the partial case $N=2$ of your inequality I prove the following inequality: $$\frac{\|x\|^2\cdot \|y\|^2+(x,y)^2}{\|Tx\|^2 \|Ty\|^2}\geqslant \frac2{{\rm tr}\, T^4}$$ for any self-adjoint positive definite operator $T$ and any two vectors $x,y$ in $\mathbb{R}^n$. If we denote $x=(a_1,\dots,a_n)$, $y=(b_1,\dots,b_n)$, $p=(a_1^2,\dots,a_n^2)$, $q=(b_1^2,\dots,b_n^2)$, $T^2=diag(s_1,\dots,s_n)$, $s=(s_1,\dots,s_n)$, we rewrite this as $$\left(\sum_{i=1}^n a_i^2\right)\left(\sum_{i=1}^n b_i^2\right)+\left(\sum_{i=1}^na_i b_i\right)^2\ge 2\frac{(p,s)\cdot (q,s)}{(s,s)},$$ and maximizing over $s$ gives you $$\sqrt{\left(\sum_{i=1}^n a_i^4\right)\left(\sum_{i=1}^n b_i^4\right)}+\sum_{i=1}^na_i^2b_i^2$$ in RHS, this is the equality case in the lemma in the cited answer.</p> Answer 12: <p>Denote the power-sum (symmetric) polynomials by $p_k(a;n)=\sum_{i=1}^na_i^k$; where $a=(a_1,\dots, a_n)$ and also we need the Hadamard product $ab=(a_1b_1,\dots,a_nb_n)$. Write $p_k(a)$ for $p_k(a;n)$ when there is no confusion. </p> <p><strong>Remark.</strong> Observe that $p_2(a)p_2(b)-p_2(ab)=\sum_{i\neq j}a_i^2b_j^2\geq0$. We prove $$[p_2(a)p_2(b)+p_1(ab)^2-p_2(ab)]^2-p_2(a^2)p_2(b^2)\geq0.\tag1$$ Induct on $n$. The case $n=2$ is noted as trivial. Assume true for $n$, we show for $n+1$; i.e. $$[(x^2+p_2(a))(y^2+p_2(b))+(xy+p_1(ab))^2-x^2y^2-p_2(ab)]^2-(x^4+p_2(a^2))(y^4+p_2(b^2))\geq0.\tag2$$ After expansion, the LHS of (2) is a polynomial in even powers of $x$ and $y$. It suffices to study the following coefficients (the others follow by symmetry):</p> <p>$[x^0y^0]$: the is exactly the induction assumption, hence positive.</p> <p>$[x^2y^0]$: $2p_2(b)[p_1(ab)^2+p_2(a)p_2(b)-p_2(ab)]\geq0$, by remark from above.</p> <p>$x^4y^0]$: $p_2(b)^2-p_2(b^2)\geq0$, by remark.</p> <p>$[x^2y^2]$: $6p_1(ab)^2+4p_2(a)p_2(b)-2p_2(ab)\geq0$, by remark.</p> <p>$[x^4y^2]$: $2p_2(b)\geq0$.</p> <p>$[x^0y^4]$: $p_2(a)^2-p_2(a^2)\geq0$, by remark.</p> <p>Therefore, the claim (1) is valid for all $n$.</p> <p>The intent of this method is to reveal that the inequality (1) is not as sharp as we may wish.</p>
A curious series related to the asymptotic behavior of the tetration
260,379
18
https://mathoverflow.net/questions/260379/a-curious-series-related-to-the-asymptotic-behavior-of-the-tetration
<p>The <a href="http://en.wikipedia.org/wiki/Tetration" rel="noreferrer">tetration</a> is denoted $^n a$, where $a$ is called the base and $n$ is called the height, and is defined for $n\in\mathbb N\cup\{-1,\,0\}$ by the recurrence $$ {^{-1} a} = 0, \quad {^{n+1} a} = a^{\left({^n a}\right)},\tag1$$ so that $${^0 a}=1, \quad {^1 a} = a, \quad {^2 a} = a^a, \quad {^3 a} = a^{a^a}, \, \dots \quad {^n a} = \underbrace{a^{a^{{.^{.^{.^a}}}}}}_{n\,\text{levels}},\tag2$$ where power towers are evaluated from top to bottom.</p> <p>Let $a$ be a real number in the interval $1&lt;a&lt;e^{1/e}$. It is convenient to introduce a notation $$\lambda = -W(-\ln a),\tag3$$ where $W(z)$ denotes the principal branch of the <a href="https://en.wikipedia.org/wiki/Lambert_W_function" rel="noreferrer">Lambert W-function</a>, satisfying $W(z) \, e^{W(z)}=z$. Note that the restrictions on $a$ imply $0&lt;\lambda&lt;1$. We may also observe that $\ln\ln a = - \lambda + \ln \lambda$.</p> <p>The sequence $\{^n a\}$ converges to a limit $$\lim_{n\to\infty} {^n a} = e^\lambda = \frac\lambda{\ln a}.\tag4$$ Its asymptotic behavior can be represented as $${^n a} = e^\lambda - c_{\lambda} \cdot \lambda^n + O(\lambda^{2n}),\tag5$$ where $c_{\lambda}$ is some positive coefficient depending on $\lambda$ (or, equivalently, on $a$). It is known that the coefficients in $O(\lambda^{2n})$ and all higher-order terms can be expressed in a closed form via $c_{\lambda}$. But, apparently, the dependency between $\lambda$ and $c_{\lambda}$ has no known simple representation. Here is how its graph looks: <a href="https://i.stack.imgur.com/ITqYR.png" rel="noreferrer"><img src="https://i.stack.imgur.com/ITqYR.png" alt="Dependency graph"></a> It appears to have a maximum near (or exactly at?) $\lambda = \ln 2$ that corresponds to $a=\sqrt2$. Numerical evidence suggests that $c_\lambda$ has a Taylor–Maclaurin series expansion with rational coefficients $$c_\lambda = \frac{\lambda }{1!}+\frac{\lambda ^2}{2!}-\frac{2 \lambda ^3}{3!}-\frac{11 \lambda ^4}{4!}-\frac{44 \lambda ^5}{5!}-\frac{89 \lambda ^6}{6!}-\frac{636 \lambda ^7}{7!}-\frac{615 \lambda ^8}{8!}-...\tag6$$ <em>(more numerators can be found <a href="https://gist.githubusercontent.com/VladimirReshetnikov/1ad1efe07457d17b56f801894db7e7c5/raw/d909240f31f7925b49a52df0929e6056354d2a58/TaylorMaclaurinNumerators.txt" rel="noreferrer">here</a>)</em></p> <p>I could not find a formula for the coefficients (apparently, they are not yet in the <a href="https://oeis.org/" rel="noreferrer">OEIS</a>), and I do not even have a proof that the coefficients given above are exact, so I am asking for your help with it. Is it possible to sum this series in terms of known special functions? Is it actually unimodal as is suggested by its graph? What is the exact location of its maximum?</p> <p>Related questions: <a href="https://math.stackexchange.com/questions/2001325/coefficients-of-a-formal-power-series-satisfying-expfz-1-fq-z-q">[1]</a><a href="https://mathoverflow.net/questions/57627/is-there-an-elegant-non-recursive-formula-for-these-coefficients-also-how-ca">[2]</a><a href="https://math.stackexchange.com/questions/1987944/complete-monotonicity-of-a-sequence-related-to-tetration">[3]</a><a href="https://mathoverflow.net/questions/259278/an-explicit-series-representation-for-the-analytic-tetration-with-complex-height">[4]</a><a href="https://math.stackexchange.com/questions/800862/convergence-of-power-towers">[5]</a><a href="https://math.stackexchange.com/questions/1528235/a-limit-related-to-super-root-tetration-inverse">[6]</a>.</p>
[ { "answer_id": 272157, "score": 9, "text": "<p>The coefficients of the series (6) can be found symbolically in $\\lambda$ and rational numbers. </p>\n\n<p>Consider some base $b$ with $1 \\le b \\le e^{1/e}$ , the fixpoint $t$ for $f(x)=b^x $ and $f(t)=t$. For bases $b$ in this interval we have a real fixpoint $t$ with $1 \\le t \\le e$ and $\\lambda = \\log(t) $ with $0 \\le \\lambda \\le 1$ with $b^t = t$ <em>(which is also attracting when $\\lambda,t,b$ are not on the bounds of the resp. intervals)</em>. </p>\n\n<p>$\\ $ </p>\n\n<p>We assume now a conjugate function $d(z)=z/t-1$ and $d°^{-1}(z)=(z+1)\\cdot t$ which I write here for shortness in sup-suffix notation $z´ = d(z) $ and $z`=d°^{-1}(z)$ .</p>\n\n<p>The <strong><em><a href=\"https://en.wikipedia.org/wiki/Schr%C3%B6der%27s_equation\" rel=\"nofollow noreferrer\">Schröder-mechanism</a></em></strong>, using this conjugacy, allows an invertible \"<em>schlicht</em>\" function $\\sigma(x)$ such that $$ f°^h(x) = \\left(\\sigma°^{-1}( \\lambda^h \\cdot \\sigma(x´))\\right)`$$ </p>\n\n<p>Now the given problem assumes $x=1$ such that $x´ = 1/t-1 = \\exp(-\\lambda)-1$ and we have explicitely \n $$ f°^h(1) = \\left(\\sigma°^{-1}( \\lambda^h \\cdot \\sigma(\\exp(-\\lambda)-1))\\right) \\cdot \\exp(\\lambda) + t$$ and because $\\sigma(x) = x + O(x^2) $ (meaning to be called \"<em>schlicht</em>\" ) also its formal power series inverse is of this form and we get\n$$ f°^h(1) = t + \\lambda^h c_\\lambda + O(\\lambda^{2h})$$ where $c_\\lambda = \\sigma(1/t-1) $<br>\nInterestingly, leaving $\\lambda$ indeterminate and $t=\\exp(\\lambda)$ unevaluated as formal power series Pari/GP gives $c_\\lambda$ as formal power series in the form as the OP in def. (6) provides its leading coefficients. </p>\n\n<hr>\n\n<blockquote>\n <p><strong>Derivation/implementation in Pari/GP</strong> </p>\n</blockquote>\n\n<p>Here I show, illustrated by sample calculations with Pari/GP, that there is no numerical approximation needed to get the coefficients at $\\lambda^k$ in definition (6). Instead there is a derivation on formal power series which provides exact rational coefficients. The method below reproduces perfectly your finding. </p>\n\n<p><hr>\n<strong>Preliminaries:</strong> a <a href=\"https://en.wikipedia.org/wiki/Carleman_matrix\" rel=\"nofollow noreferrer\">Carleman-matrix</a> <strong><em>B</em></strong> associated to a function $f(x)$ which has a power series with more than zero radius of convergence (and generalized: can be Euler-summed in a certain interval of its argument) is initially only a collection of column-vectors $B[,c]$ <em>($c$ indexing the c-olumn)</em>, which contain the coefficients of the formal power series of $f(x)$ and more precisely of $f(x)^c$ for $c=0 \\ldots \\infty$ in the column of index $c$ , and is thus in both direction of infinite size. It is an important special case when the power series of some function $f(x)$ has no constant term: then the shape of $B$ is lower triangular and the analytic handling of such Carlemanmatrices has been substantially established. We'll need this in the derivation below.<br>\n<sub> It is not trivial to show, that such a <em>collection</em> of columns-vectors can indeed be used as <strong><em>Matrix</em></strong>, applicable to concepts of matrix-algebra and -analysis, for instance raising to powers and especially diagonalization - but there is reliable literature about this so that we can omit such a step here. One pioneer is surely Eri Jabotinsky, see for instance this 1963-article <a href=\"http://www.jstor.org/stable/1993593\" rel=\"nofollow noreferrer\">\"Analytic iteration\"</a> (introducing the term \"representation matrix\")</sub> </p>\n\n<p>After the Carlemanmatrix <strong><em>B</em></strong> contains the coefficients of formal power series we need also a construct $V(x)$ as row vector $V(x)=[1,x,x^2,x^3,\\ldots]$ which I call of \"Vandermonde\"-type here. Using $V(x)$ as diagonal-matrix I pre-superscribe this like $ \\;^dV(x)$ \n<hr>\nFor instance the Carlemanmatrix $B_e$ for the function $f(x) = \\exp(x)$ looks like</p>\n\n<pre><code> 1 1 1 1 1 ...\n 0 1 2 3 4 ...\n 0 1/2! 2^2/2! 3^2/2! 4^2/2! ...\n 0 1/3! 2^3/3! 3^3/3! 4^3/3! ...\n 0 1/4! 2^4/4! 3^4/4! 4^4/4! ...\n ... ... ... ... .... ... \n</code></pre>\n\n<p>The clue with such Carlemanmatrices is, that the left-multiplication of some $B$ with a vector $V(x)$ as in\n $$V(x) \\cdot B = V(f(x)) \\tag 1$$ \ngives a resulting vector of the same structure $V(f(x)) = [ 1 , f(x), f(x)^2, f(x)^3, ... ]$ .<br>\nOf course, using the only \"relevant\" column in $B$ alone , then this gives the scalar expression $f(x)$: \n $$V(x) \\cdot B [,1] = f(x) \\tag {1.1}$$ </p>\n\n<p>This means: a composition of a function with itself $f(f(x))$ (in terms of its power series) is mapped to a <em>repeated multiplication</em> with its Carleman-matrix or in other words: is mapped to multiplication with <strong><em>powers</em></strong> of its Carlemanmatrix. </p>\n\n<p>The further clue is now, that for appropriate matrices a <strong><em>fractional</em></strong> power can be determined and for that cases we can thus easily define power series for the fractional iterates of the associated function. </p>\n\n<hr> \n\n<p><strong>Technically</strong> this means that the concept of diagonalization can be transferred to that Carleman-ansatz (of course for suitable Carlemanmatrices $B$). A case, for which this is widely accepted is the case where the Carleman-matrices for some function $f(x)$ is lower triangular.</p>\n\n<p>We have the function $f(x) = b^x $ and by the definition in the OP $1 \\lt b \\lt e^{1/e}$ . With this restriction, $f(x)$ has a fixpoint $t$ such that $f(t)=t$ (or: $b^t=t$ ) and also this fixpoint is attracting (which is important here). We can, as I wrote in the first part of text above, apply conjugacy to arrive at the function $g(x)=f(x`)´ = t^x-1 = ux + (ux)^2/2! + O(x^3) $ (for notational convenience I write $u$ instead of $\\lambda$ in the following). </p>\n\n<p>Given the lower triangular matrix of Stirlingnumbers $2$'nd kind $S2$ and its factorially similarity-scaled transform which I denote here as $fS2F$ we have the lower triangular Carlemanmatrix $U_t$ for $g(x)$ by $$ U_t = \\; ^dV(u) \\cdot fS2F \\tag {2.1}$$</p>\n\n<hr> \n\n<p>The top-left edge of $U_t$ looks like</p>\n\n<pre><code> 1 . . . . . \n 0 u/1! . . . . \n 0 u^2/2! u^2*2!/2! . . . \n 0 u^3/3! 3*u^3*2!/3! u^3*3!/3! . . \n 0 u^4/4! 7*u^4*2!/4! 6*u^4*3!/4! u^4*4!/4! . \n 0 u^5/5! 15*u^5*2!/5! 25*u^5*3!/5! 10*u^5*4!/5! u^5*5!/5! \n</code></pre>\n\n<p>or with cancelled entries </p>\n\n<pre><code> 1 . . . . .\n 0 u . . . .\n 0 1/2*u^2 u^2 . . .\n 0 1/6*u^3 u^3 u^3 . .\n 0 1/24*u^4 7/12*u^4 3/2*u^4 u^4 .\n 0 1/120*u^5 1/4*u^5 5/4*u^5 2*u^5 u^5\n</code></pre>\n\n<p>and taking the coefficients from the second column $$ g(x) = \\sum_{k=1}^\\infty u^k \\cdot \\frac{x^k}{ k!} \\tag {2.2}$$ \n<hr>\nThe formal power series for $h$'th fractional iterates $g°^h(x)$ are now generated by fractional powers of $U_t$ instead and this can be done by diagonalization: </p>\n\n<p>$$ U_t = M \\cdot D \\cdot W \\qquad \\qquad \\text{where I denote $W$ for } M^{-1} \\tag {3.1}$$\nwhere also the diagonal entries in $D$ are the consecutive powers of $u$\n$$ D = \\; ^d V(u) \\text{ and } D^h = \\; ^dV(u^h) \\tag {3.2}$$\nsuch that\n$$ U_t^h = M \\cdot \\; ^dV(u^h) \\cdot W \\tag {3.3} $$</p>\n\n<p>Unfortunately the current Pari/GP-version's diagonalization procedure cannot exploit the advantage of $U_t$ being triangular and also cannot keep the parameter $u$ symbolically. But it is not difficult to have a small recursive routine to generate $M$ , $D$ and $W (=M^{-1})$ for the triangular Carleman-case; the additional norming of $M$ to have a unit-diagonal is also included and with this the matrix $M$ is also of the Carleman-type (for details and an implementation in Pari/GP see the explanation in my linked article). The matrix <strong><em>M</em></strong> is now the Carlemanmatrix for the above introduced Schröder-function $\\sigma(x)$ which takes its coefficients from the second column in <strong><em>M</em></strong>. </p>\n\n<p>The top-left of the matrix $M$ looks like</p>\n\n<pre><code> 1 . . .\n 0 1 . .\n 0 u/(-2*u+2) 1 .\n 0 (2*u^3+u^2)/(6*u^3-6*u^2-6*u+6) u/(-u+1) 1\n</code></pre>\n\n<p>The entries in the (only relevant) $2$'nd column have a partially discernable pattern; I've described this in earlier discussions, see for instance <a href=\"http://go.helms-net.de/math/tetdocs/APT.htm\" rel=\"nofollow noreferrer\">here</a>, but which is too much to show at this place.<br>\n<hr>\nHaving this we can write for the $h$'th fractional iterate $g(x)$ of some $x$: </p>\n\n<p>$$ \\begin{array}{rll} V(x) \\cdot U_t^h &amp;= V( g°^h(x)) \\\\ &amp;= V(x) \\cdot (M \\cdot \\;^dV(u^h) \\cdot W) \\qquad \\qquad \\text{ and by associativity}\\\\\n &amp;= (V(x) \\cdot M) \\cdot \\;^dV(u^h) \\cdot W \\end{array} \\tag {4.1}$$</p>\n\n<p>Because $M$ is of Carlemantype we have also\n $$ V(x) \\cdot M = V(\\sigma(x)) \\tag {4.2}$$ \nor -using also conjugacy- $$ V(x´) \\cdot M = V(\\sigma(x´)) \\tag {4.3}$$\n and it is relevant here, that $\\sigma(x´)$ is <strong><em>identical to the (normed) Schröder-function</em></strong> for the iteration of $b^x$. </p>\n\n<p>In the problem as given in the OP we assume $x=1$ or $x´=1/t-1 = \\exp(-u)-1$ and to have the Schröder-function symbolically we have to look at\n$$ V(1´) \\cdot M = V(\\exp(-u)-1) \\cdot M = V(\\sigma(\\exp(-u)-1)) \\tag {4.4} $$ and this can still be done symbolically in $u$; having <strong><em>M</em></strong> in its symbolical representation, Pari/GP is able to give the leading part of $\\sigma(\\exp(-u)-1)$ by formal expansion of $\\exp(-u)$ as power series taken as argument of $\\sigma(\\cdot)$ with still rational coefficients. </p>\n\n<p>Having the left part of $(4.2)$ evaluated we need now\n$$ V(\\sigma(\\exp(-u)-1)) \\cdot \\; ^dV(u^h) \\cdot W[,1] = y =\\sigma°^{-1}(u^h \\cdot \\sigma(1´))\\tag {4.5}$$\nby the relevant $2$'nd column of $W$ only. </p>\n\n<p>Note that this gives a power series in $u^h$ with the (constant and rational) coefficients in $W[,1]$ and powers of the expression $\\sigma(\\exp(-u)-1)$ . To help Pari/GP to keep $u^h$ as a single symbol, we need to substitute it by the name of a unevaluated variable, say $u_h$ or $v$ and re-substitute it later after series-expansions.</p>\n\n<p>The final result $f°^h(1) $ occurs by the inverse conjugacy on $y$\n$$f°^h(1) = y` = (y+1)\\cdot t = y\\cdot t + t =t + y \\cdot \\exp(u) \\tag {4.6} $$\nwhich can still be expressed as formal power series in the indeterminate $u$ and $u^h$. </p>\n\n<p>The OP asks now for the first coefficient in $y \\cdot \\exp(u) $ and namely for its series expression when $u$ <em>and</em> $v (=u^h) $ are both kept as indeterminate parameter. Pari/GP is still able to evaluate the last expression symbolically and arrives at the same set of coefficients as given in eq (6) in the OP.\n<hr>\n<strong>Additional computation:</strong> a short recalculation gives also the leading terms of the powerseries of the following coefficients in (5) of the OP, which I index here as $c_{\\lambda,k}$ : </p>\n\n<p>$$\\small \\begin{array} {r}\nc_{\\lambda,1}=&amp; - 1 u^1/1!&amp; - 1 u^2/2!&amp; + 2 u^3/3!&amp; + 11 u^4/4!&amp; + 44 u^5/5!&amp; + 89 u^6/6!&amp; + 636 u^7/7!&amp; + O(u^8) \\\\\nc_{\\lambda,2}=&amp;&amp;&amp; - 3 u^3/3!&amp; - 12 u^4/4!&amp; - 5 u^5/5!&amp; + 150 u^6/6!&amp; + 1463 u^7/7!&amp; + O(u^8) \\\\\nc_{\\lambda,3}=&amp;&amp;&amp;&amp;&amp; - 40 u^5/5!&amp; - 240 u^6/6!&amp; - 840 u^7/7!&amp; + O(u^8) \\\\\nc_{\\lambda,4}=&amp;&amp;&amp;&amp;&amp;&amp;&amp; - 1260 u^7/7!&amp; + O(u^8) \\\\\n \\end{array}\n$$\nand recalling that $\\; ^hb $ is my notation for your $ \\; ^na $ we get\n$$ \\; ^na = \\; ^hb = t +c_{\\lambda,1} u^h + c_{\\lambda,2} u^{2h} + c_{\\lambda,3} u^{3h} + \\cdots \\tag 5$$\nRemark: I don't think that a closing term $O(u^{4h})$ makes really sense here because we assume $h&gt;-2$ to be of <em>any real value</em>, especially fractional, and I'm not firm with using that notation in such cases.</p>\n" }, { "answer_id": 272150, "score": 5, "text": "<p>Let $$c_n(\\lambda)=\\frac{e^\\lambda-{^n a}}{ \\lambda^n},\\quad d_n(\\lambda)=e^{-\\lambda}c_n(\\lambda)$$ ($d_n$ seems to be simpler). In particular initial functions $c_0(\\lambda)=e^\\lambda-1$ and $d_0(\\lambda)=1-e^{-\\lambda}\\sim\\lambda$ (as $\\lambda\\to 0$) have good power series of the variable $\\lambda$. We are intersted in the limits $$c_\\lambda=\\lim\\limits_{n\\to\\infty}c_n(\\lambda),\\quad d_\\lambda=\\lim\\limits_{n\\to\\infty}d_n(\\lambda).$$ The recurrence\n\\begin{gather*}\nd_{n+1}(\\lambda)=\\frac{1-e^{-\\lambda^{n+1}d_n(\\lambda)}}{\\lambda^{n+1} }\n\\end{gather*}\nfollows from the definition. By induction $d_n(\\lambda)\\sim\\lambda$. Expanding the recurrence above we get\n\\begin{gather*}\nd_{n+1}(\\lambda)=d_n(\\lambda)-\\frac{\\lambda^{n+1}d_n^2(\\lambda)}{2}+\\ldots\\equiv d_n(\\lambda)\\pmod{\\lambda^{n+3} }.\n\\end{gather*}\nIt means that the limit $d_\\lambda$ (as a power series) does exist.</p>\n\n<p><strong>EDT.</strong> Numerators $a(k)$ of $c_\\lambda$-coefficients are not very random. At least first $41$ coefficients from OP satisfy the congruences\n\\begin{gather*}\na(k)\\equiv\\begin{cases}\n0,&amp;\\text{if } k\\equiv 1\\pmod{2 },\\\\\n1,&amp;\\text{if }k\\equiv 0\\pmod{4 },\\\\\n3,&amp;\\text{if }k\\equiv 2\\pmod{4 },\\\\\n\\end{cases}\\pmod{4 }\\quad(k\\ge 4); \\\\\na(k)\\equiv\\begin{cases}\n0,&amp;\\text{if }k\\equiv 1,2\\pmod{6 },\\\\\n1,&amp;\\text{if }k\\equiv0,3,4,5\\pmod{6 },\\\\\n\\end{cases}\\pmod{3 }\\quad(k\\ge 3).\n\\end{gather*}</p>\n" }, { "answer_id": 272338, "score": 2, "text": "<p><strong><em>This is not an answer, only an extended comment at V. Reshetnikov's request in his comment <a href=\"https://mathoverflow.net/a/272157/7710\">at my answer</a></em></strong> </p>\n\n<p>@VladimirReshetnikov - not so far now. I think this problem of a closed form\nis also intimately related to the so far unsolved question of the \"mag\"-numbers as discussed in \n<a href=\"https://mathoverflow.net/questions/57627\">\"Is there an “elegant” non-recursive formula for these coefficients?\"</a> \nwhich is focusing the series expansion for the Schröder-function $\\sigma()$ as I've used it here.</p>\n\n<p>Also your approach using q-binomials for a series <a href=\"https://mathoverflow.net/questions/259278\">\"An explicit series representation for the analytic tetration with complex height\"</a> was/is much interesting for me - but although I had similar analyses involving q-binomials I didn't arrive at any conciser formulae than the formal algebraic descriptions following from the matrix-based concepts as discussed, for instance, here: <a href=\"http://go.helms-net.de/math/tetdocs/ExponentialPolynomialInterpolation.pdf\" rel=\"nofollow noreferrer\">\"Exponential polynomial interpolation\"</a> and also here <a href=\"http://go.helms-net.de/math/tetdocs/BinomialDiagonalization.htm\" rel=\"nofollow noreferrer\">\"Relation between binomial- and diagonalization method\"</a> but at least showing at its end the series <strong><em>(5)</em></strong> in your OP assuming $\\lambda=\\log 2$ with evaluated coefficients in real numbers and which gives also a short resume of the -as I christened it- \"exponential polynomial interpolation\" method which involves the q-binomials.</p>\n\n<p>In general: the -in my opinion- most sophisticated explication of the coefficients of the function $t^x-1$ (and by conjugacy as well of $b^x$ ) in terms of the logarithm $u$ of $t$ and of $u^h$ comes via the diagonalization-approach as shown in <a href=\"http://go.helms-net.de/math/tetdocs/APT.htm\" rel=\"nofollow noreferrer\">\"Eigendecomposition of triangular matrix-operators (here: Ut)\"</a> from where the coefficients of your formulae <strong><em>(5)</em></strong> and <strong><em>(6)</em></strong> can be derived, and even symbolically<br>\nBut as well as with the q-binomial approach I did not yet find any simplifications/closed forms besides that (recursive)\nbasic generating schemes for that formulae/coefficients, I am really sorry. </p>\n\n<p>Of course, in case I'll find something worth sometime I'll surely link to that discussion here.</p>\n" }, { "answer_id": 272157, "score": 9, "text": "<p>The coefficients of the series (6) can be found symbolically in $\\lambda$ and rational numbers. </p>\n\n<p>Consider some base $b$ with $1 \\le b \\le e^{1/e}$ , the fixpoint $t$ for $f(x)=b^x $ and $f(t)=t$. For bases $b$ in this interval we have a real fixpoint $t$ with $1 \\le t \\le e$ and $\\lambda = \\log(t) $ with $0 \\le \\lambda \\le 1$ with $b^t = t$ <em>(which is also attracting when $\\lambda,t,b$ are not on the bounds of the resp. intervals)</em>. </p>\n\n<p>$\\ $ </p>\n\n<p>We assume now a conjugate function $d(z)=z/t-1$ and $d°^{-1}(z)=(z+1)\\cdot t$ which I write here for shortness in sup-suffix notation $z´ = d(z) $ and $z`=d°^{-1}(z)$ .</p>\n\n<p>The <strong><em><a href=\"https://en.wikipedia.org/wiki/Schr%C3%B6der%27s_equation\" rel=\"nofollow noreferrer\">Schröder-mechanism</a></em></strong>, using this conjugacy, allows an invertible \"<em>schlicht</em>\" function $\\sigma(x)$ such that $$ f°^h(x) = \\left(\\sigma°^{-1}( \\lambda^h \\cdot \\sigma(x´))\\right)`$$ </p>\n\n<p>Now the given problem assumes $x=1$ such that $x´ = 1/t-1 = \\exp(-\\lambda)-1$ and we have explicitely \n $$ f°^h(1) = \\left(\\sigma°^{-1}( \\lambda^h \\cdot \\sigma(\\exp(-\\lambda)-1))\\right) \\cdot \\exp(\\lambda) + t$$ and because $\\sigma(x) = x + O(x^2) $ (meaning to be called \"<em>schlicht</em>\" ) also its formal power series inverse is of this form and we get\n$$ f°^h(1) = t + \\lambda^h c_\\lambda + O(\\lambda^{2h})$$ where $c_\\lambda = \\sigma(1/t-1) $<br>\nInterestingly, leaving $\\lambda$ indeterminate and $t=\\exp(\\lambda)$ unevaluated as formal power series Pari/GP gives $c_\\lambda$ as formal power series in the form as the OP in def. (6) provides its leading coefficients. </p>\n\n<hr>\n\n<blockquote>\n <p><strong>Derivation/implementation in Pari/GP</strong> </p>\n</blockquote>\n\n<p>Here I show, illustrated by sample calculations with Pari/GP, that there is no numerical approximation needed to get the coefficients at $\\lambda^k$ in definition (6). Instead there is a derivation on formal power series which provides exact rational coefficients. The method below reproduces perfectly your finding. </p>\n\n<p><hr>\n<strong>Preliminaries:</strong> a <a href=\"https://en.wikipedia.org/wiki/Carleman_matrix\" rel=\"nofollow noreferrer\">Carleman-matrix</a> <strong><em>B</em></strong> associated to a function $f(x)$ which has a power series with more than zero radius of convergence (and generalized: can be Euler-summed in a certain interval of its argument) is initially only a collection of column-vectors $B[,c]$ <em>($c$ indexing the c-olumn)</em>, which contain the coefficients of the formal power series of $f(x)$ and more precisely of $f(x)^c$ for $c=0 \\ldots \\infty$ in the column of index $c$ , and is thus in both direction of infinite size. It is an important special case when the power series of some function $f(x)$ has no constant term: then the shape of $B$ is lower triangular and the analytic handling of such Carlemanmatrices has been substantially established. We'll need this in the derivation below.<br>\n<sub> It is not trivial to show, that such a <em>collection</em> of columns-vectors can indeed be used as <strong><em>Matrix</em></strong>, applicable to concepts of matrix-algebra and -analysis, for instance raising to powers and especially diagonalization - but there is reliable literature about this so that we can omit such a step here. One pioneer is surely Eri Jabotinsky, see for instance this 1963-article <a href=\"http://www.jstor.org/stable/1993593\" rel=\"nofollow noreferrer\">\"Analytic iteration\"</a> (introducing the term \"representation matrix\")</sub> </p>\n\n<p>After the Carlemanmatrix <strong><em>B</em></strong> contains the coefficients of formal power series we need also a construct $V(x)$ as row vector $V(x)=[1,x,x^2,x^3,\\ldots]$ which I call of \"Vandermonde\"-type here. Using $V(x)$ as diagonal-matrix I pre-superscribe this like $ \\;^dV(x)$ \n<hr>\nFor instance the Carlemanmatrix $B_e$ for the function $f(x) = \\exp(x)$ looks like</p>\n\n<pre><code> 1 1 1 1 1 ...\n 0 1 2 3 4 ...\n 0 1/2! 2^2/2! 3^2/2! 4^2/2! ...\n 0 1/3! 2^3/3! 3^3/3! 4^3/3! ...\n 0 1/4! 2^4/4! 3^4/4! 4^4/4! ...\n ... ... ... ... .... ... \n</code></pre>\n\n<p>The clue with such Carlemanmatrices is, that the left-multiplication of some $B$ with a vector $V(x)$ as in\n $$V(x) \\cdot B = V(f(x)) \\tag 1$$ \ngives a resulting vector of the same structure $V(f(x)) = [ 1 , f(x), f(x)^2, f(x)^3, ... ]$ .<br>\nOf course, using the only \"relevant\" column in $B$ alone , then this gives the scalar expression $f(x)$: \n $$V(x) \\cdot B [,1] = f(x) \\tag {1.1}$$ </p>\n\n<p>This means: a composition of a function with itself $f(f(x))$ (in terms of its power series) is mapped to a <em>repeated multiplication</em> with its Carleman-matrix or in other words: is mapped to multiplication with <strong><em>powers</em></strong> of its Carlemanmatrix. </p>\n\n<p>The further clue is now, that for appropriate matrices a <strong><em>fractional</em></strong> power can be determined and for that cases we can thus easily define power series for the fractional iterates of the associated function. </p>\n\n<hr> \n\n<p><strong>Technically</strong> this means that the concept of diagonalization can be transferred to that Carleman-ansatz (of course for suitable Carlemanmatrices $B$). A case, for which this is widely accepted is the case where the Carleman-matrices for some function $f(x)$ is lower triangular.</p>\n\n<p>We have the function $f(x) = b^x $ and by the definition in the OP $1 \\lt b \\lt e^{1/e}$ . With this restriction, $f(x)$ has a fixpoint $t$ such that $f(t)=t$ (or: $b^t=t$ ) and also this fixpoint is attracting (which is important here). We can, as I wrote in the first part of text above, apply conjugacy to arrive at the function $g(x)=f(x`)´ = t^x-1 = ux + (ux)^2/2! + O(x^3) $ (for notational convenience I write $u$ instead of $\\lambda$ in the following). </p>\n\n<p>Given the lower triangular matrix of Stirlingnumbers $2$'nd kind $S2$ and its factorially similarity-scaled transform which I denote here as $fS2F$ we have the lower triangular Carlemanmatrix $U_t$ for $g(x)$ by $$ U_t = \\; ^dV(u) \\cdot fS2F \\tag {2.1}$$</p>\n\n<hr> \n\n<p>The top-left edge of $U_t$ looks like</p>\n\n<pre><code> 1 . . . . . \n 0 u/1! . . . . \n 0 u^2/2! u^2*2!/2! . . . \n 0 u^3/3! 3*u^3*2!/3! u^3*3!/3! . . \n 0 u^4/4! 7*u^4*2!/4! 6*u^4*3!/4! u^4*4!/4! . \n 0 u^5/5! 15*u^5*2!/5! 25*u^5*3!/5! 10*u^5*4!/5! u^5*5!/5! \n</code></pre>\n\n<p>or with cancelled entries </p>\n\n<pre><code> 1 . . . . .\n 0 u . . . .\n 0 1/2*u^2 u^2 . . .\n 0 1/6*u^3 u^3 u^3 . .\n 0 1/24*u^4 7/12*u^4 3/2*u^4 u^4 .\n 0 1/120*u^5 1/4*u^5 5/4*u^5 2*u^5 u^5\n</code></pre>\n\n<p>and taking the coefficients from the second column $$ g(x) = \\sum_{k=1}^\\infty u^k \\cdot \\frac{x^k}{ k!} \\tag {2.2}$$ \n<hr>\nThe formal power series for $h$'th fractional iterates $g°^h(x)$ are now generated by fractional powers of $U_t$ instead and this can be done by diagonalization: </p>\n\n<p>$$ U_t = M \\cdot D \\cdot W \\qquad \\qquad \\text{where I denote $W$ for } M^{-1} \\tag {3.1}$$\nwhere also the diagonal entries in $D$ are the consecutive powers of $u$\n$$ D = \\; ^d V(u) \\text{ and } D^h = \\; ^dV(u^h) \\tag {3.2}$$\nsuch that\n$$ U_t^h = M \\cdot \\; ^dV(u^h) \\cdot W \\tag {3.3} $$</p>\n\n<p>Unfortunately the current Pari/GP-version's diagonalization procedure cannot exploit the advantage of $U_t$ being triangular and also cannot keep the parameter $u$ symbolically. But it is not difficult to have a small recursive routine to generate $M$ , $D$ and $W (=M^{-1})$ for the triangular Carleman-case; the additional norming of $M$ to have a unit-diagonal is also included and with this the matrix $M$ is also of the Carleman-type (for details and an implementation in Pari/GP see the explanation in my linked article). The matrix <strong><em>M</em></strong> is now the Carlemanmatrix for the above introduced Schröder-function $\\sigma(x)$ which takes its coefficients from the second column in <strong><em>M</em></strong>. </p>\n\n<p>The top-left of the matrix $M$ looks like</p>\n\n<pre><code> 1 . . .\n 0 1 . .\n 0 u/(-2*u+2) 1 .\n 0 (2*u^3+u^2)/(6*u^3-6*u^2-6*u+6) u/(-u+1) 1\n</code></pre>\n\n<p>The entries in the (only relevant) $2$'nd column have a partially discernable pattern; I've described this in earlier discussions, see for instance <a href=\"http://go.helms-net.de/math/tetdocs/APT.htm\" rel=\"nofollow noreferrer\">here</a>, but which is too much to show at this place.<br>\n<hr>\nHaving this we can write for the $h$'th fractional iterate $g(x)$ of some $x$: </p>\n\n<p>$$ \\begin{array}{rll} V(x) \\cdot U_t^h &amp;= V( g°^h(x)) \\\\ &amp;= V(x) \\cdot (M \\cdot \\;^dV(u^h) \\cdot W) \\qquad \\qquad \\text{ and by associativity}\\\\\n &amp;= (V(x) \\cdot M) \\cdot \\;^dV(u^h) \\cdot W \\end{array} \\tag {4.1}$$</p>\n\n<p>Because $M$ is of Carlemantype we have also\n $$ V(x) \\cdot M = V(\\sigma(x)) \\tag {4.2}$$ \nor -using also conjugacy- $$ V(x´) \\cdot M = V(\\sigma(x´)) \\tag {4.3}$$\n and it is relevant here, that $\\sigma(x´)$ is <strong><em>identical to the (normed) Schröder-function</em></strong> for the iteration of $b^x$. </p>\n\n<p>In the problem as given in the OP we assume $x=1$ or $x´=1/t-1 = \\exp(-u)-1$ and to have the Schröder-function symbolically we have to look at\n$$ V(1´) \\cdot M = V(\\exp(-u)-1) \\cdot M = V(\\sigma(\\exp(-u)-1)) \\tag {4.4} $$ and this can still be done symbolically in $u$; having <strong><em>M</em></strong> in its symbolical representation, Pari/GP is able to give the leading part of $\\sigma(\\exp(-u)-1)$ by formal expansion of $\\exp(-u)$ as power series taken as argument of $\\sigma(\\cdot)$ with still rational coefficients. </p>\n\n<p>Having the left part of $(4.2)$ evaluated we need now\n$$ V(\\sigma(\\exp(-u)-1)) \\cdot \\; ^dV(u^h) \\cdot W[,1] = y =\\sigma°^{-1}(u^h \\cdot \\sigma(1´))\\tag {4.5}$$\nby the relevant $2$'nd column of $W$ only. </p>\n\n<p>Note that this gives a power series in $u^h$ with the (constant and rational) coefficients in $W[,1]$ and powers of the expression $\\sigma(\\exp(-u)-1)$ . To help Pari/GP to keep $u^h$ as a single symbol, we need to substitute it by the name of a unevaluated variable, say $u_h$ or $v$ and re-substitute it later after series-expansions.</p>\n\n<p>The final result $f°^h(1) $ occurs by the inverse conjugacy on $y$\n$$f°^h(1) = y` = (y+1)\\cdot t = y\\cdot t + t =t + y \\cdot \\exp(u) \\tag {4.6} $$\nwhich can still be expressed as formal power series in the indeterminate $u$ and $u^h$. </p>\n\n<p>The OP asks now for the first coefficient in $y \\cdot \\exp(u) $ and namely for its series expression when $u$ <em>and</em> $v (=u^h) $ are both kept as indeterminate parameter. Pari/GP is still able to evaluate the last expression symbolically and arrives at the same set of coefficients as given in eq (6) in the OP.\n<hr>\n<strong>Additional computation:</strong> a short recalculation gives also the leading terms of the powerseries of the following coefficients in (5) of the OP, which I index here as $c_{\\lambda,k}$ : </p>\n\n<p>$$\\small \\begin{array} {r}\nc_{\\lambda,1}=&amp; - 1 u^1/1!&amp; - 1 u^2/2!&amp; + 2 u^3/3!&amp; + 11 u^4/4!&amp; + 44 u^5/5!&amp; + 89 u^6/6!&amp; + 636 u^7/7!&amp; + O(u^8) \\\\\nc_{\\lambda,2}=&amp;&amp;&amp; - 3 u^3/3!&amp; - 12 u^4/4!&amp; - 5 u^5/5!&amp; + 150 u^6/6!&amp; + 1463 u^7/7!&amp; + O(u^8) \\\\\nc_{\\lambda,3}=&amp;&amp;&amp;&amp;&amp; - 40 u^5/5!&amp; - 240 u^6/6!&amp; - 840 u^7/7!&amp; + O(u^8) \\\\\nc_{\\lambda,4}=&amp;&amp;&amp;&amp;&amp;&amp;&amp; - 1260 u^7/7!&amp; + O(u^8) \\\\\n \\end{array}\n$$\nand recalling that $\\; ^hb $ is my notation for your $ \\; ^na $ we get\n$$ \\; ^na = \\; ^hb = t +c_{\\lambda,1} u^h + c_{\\lambda,2} u^{2h} + c_{\\lambda,3} u^{3h} + \\cdots \\tag 5$$\nRemark: I don't think that a closing term $O(u^{4h})$ makes really sense here because we assume $h&gt;-2$ to be of <em>any real value</em>, especially fractional, and I'm not firm with using that notation in such cases.</p>\n" }, { "answer_id": 272150, "score": 5, "text": "<p>Let $$c_n(\\lambda)=\\frac{e^\\lambda-{^n a}}{ \\lambda^n},\\quad d_n(\\lambda)=e^{-\\lambda}c_n(\\lambda)$$ ($d_n$ seems to be simpler). In particular initial functions $c_0(\\lambda)=e^\\lambda-1$ and $d_0(\\lambda)=1-e^{-\\lambda}\\sim\\lambda$ (as $\\lambda\\to 0$) have good power series of the variable $\\lambda$. We are intersted in the limits $$c_\\lambda=\\lim\\limits_{n\\to\\infty}c_n(\\lambda),\\quad d_\\lambda=\\lim\\limits_{n\\to\\infty}d_n(\\lambda).$$ The recurrence\n\\begin{gather*}\nd_{n+1}(\\lambda)=\\frac{1-e^{-\\lambda^{n+1}d_n(\\lambda)}}{\\lambda^{n+1} }\n\\end{gather*}\nfollows from the definition. By induction $d_n(\\lambda)\\sim\\lambda$. Expanding the recurrence above we get\n\\begin{gather*}\nd_{n+1}(\\lambda)=d_n(\\lambda)-\\frac{\\lambda^{n+1}d_n^2(\\lambda)}{2}+\\ldots\\equiv d_n(\\lambda)\\pmod{\\lambda^{n+3} }.\n\\end{gather*}\nIt means that the limit $d_\\lambda$ (as a power series) does exist.</p>\n\n<p><strong>EDT.</strong> Numerators $a(k)$ of $c_\\lambda$-coefficients are not very random. At least first $41$ coefficients from OP satisfy the congruences\n\\begin{gather*}\na(k)\\equiv\\begin{cases}\n0,&amp;\\text{if } k\\equiv 1\\pmod{2 },\\\\\n1,&amp;\\text{if }k\\equiv 0\\pmod{4 },\\\\\n3,&amp;\\text{if }k\\equiv 2\\pmod{4 },\\\\\n\\end{cases}\\pmod{4 }\\quad(k\\ge 4); \\\\\na(k)\\equiv\\begin{cases}\n0,&amp;\\text{if }k\\equiv 1,2\\pmod{6 },\\\\\n1,&amp;\\text{if }k\\equiv0,3,4,5\\pmod{6 },\\\\\n\\end{cases}\\pmod{3 }\\quad(k\\ge 3).\n\\end{gather*}</p>\n" }, { "answer_id": 272338, "score": 2, "text": "<p><strong><em>This is not an answer, only an extended comment at V. Reshetnikov's request in his comment <a href=\"https://mathoverflow.net/a/272157/7710\">at my answer</a></em></strong> </p>\n\n<p>@VladimirReshetnikov - not so far now. I think this problem of a closed form\nis also intimately related to the so far unsolved question of the \"mag\"-numbers as discussed in \n<a href=\"https://mathoverflow.net/questions/57627\">\"Is there an “elegant” non-recursive formula for these coefficients?\"</a> \nwhich is focusing the series expansion for the Schröder-function $\\sigma()$ as I've used it here.</p>\n\n<p>Also your approach using q-binomials for a series <a href=\"https://mathoverflow.net/questions/259278\">\"An explicit series representation for the analytic tetration with complex height\"</a> was/is much interesting for me - but although I had similar analyses involving q-binomials I didn't arrive at any conciser formulae than the formal algebraic descriptions following from the matrix-based concepts as discussed, for instance, here: <a href=\"http://go.helms-net.de/math/tetdocs/ExponentialPolynomialInterpolation.pdf\" rel=\"nofollow noreferrer\">\"Exponential polynomial interpolation\"</a> and also here <a href=\"http://go.helms-net.de/math/tetdocs/BinomialDiagonalization.htm\" rel=\"nofollow noreferrer\">\"Relation between binomial- and diagonalization method\"</a> but at least showing at its end the series <strong><em>(5)</em></strong> in your OP assuming $\\lambda=\\log 2$ with evaluated coefficients in real numbers and which gives also a short resume of the -as I christened it- \"exponential polynomial interpolation\" method which involves the q-binomials.</p>\n\n<p>In general: the -in my opinion- most sophisticated explication of the coefficients of the function $t^x-1$ (and by conjugacy as well of $b^x$ ) in terms of the logarithm $u$ of $t$ and of $u^h$ comes via the diagonalization-approach as shown in <a href=\"http://go.helms-net.de/math/tetdocs/APT.htm\" rel=\"nofollow noreferrer\">\"Eigendecomposition of triangular matrix-operators (here: Ut)\"</a> from where the coefficients of your formulae <strong><em>(5)</em></strong> and <strong><em>(6)</em></strong> can be derived, and even symbolically<br>\nBut as well as with the q-binomial approach I did not yet find any simplifications/closed forms besides that (recursive)\nbasic generating schemes for that formulae/coefficients, I am really sorry. </p>\n\n<p>Of course, in case I'll find something worth sometime I'll surely link to that discussion here.</p>\n" }, { "answer_id": 272157, "score": 9, "text": "<p>The coefficients of the series (6) can be found symbolically in $\\lambda$ and rational numbers. </p>\n\n<p>Consider some base $b$ with $1 \\le b \\le e^{1/e}$ , the fixpoint $t$ for $f(x)=b^x $ and $f(t)=t$. For bases $b$ in this interval we have a real fixpoint $t$ with $1 \\le t \\le e$ and $\\lambda = \\log(t) $ with $0 \\le \\lambda \\le 1$ with $b^t = t$ <em>(which is also attracting when $\\lambda,t,b$ are not on the bounds of the resp. intervals)</em>. </p>\n\n<p>$\\ $ </p>\n\n<p>We assume now a conjugate function $d(z)=z/t-1$ and $d°^{-1}(z)=(z+1)\\cdot t$ which I write here for shortness in sup-suffix notation $z´ = d(z) $ and $z`=d°^{-1}(z)$ .</p>\n\n<p>The <strong><em><a href=\"https://en.wikipedia.org/wiki/Schr%C3%B6der%27s_equation\" rel=\"nofollow noreferrer\">Schröder-mechanism</a></em></strong>, using this conjugacy, allows an invertible \"<em>schlicht</em>\" function $\\sigma(x)$ such that $$ f°^h(x) = \\left(\\sigma°^{-1}( \\lambda^h \\cdot \\sigma(x´))\\right)`$$ </p>\n\n<p>Now the given problem assumes $x=1$ such that $x´ = 1/t-1 = \\exp(-\\lambda)-1$ and we have explicitely \n $$ f°^h(1) = \\left(\\sigma°^{-1}( \\lambda^h \\cdot \\sigma(\\exp(-\\lambda)-1))\\right) \\cdot \\exp(\\lambda) + t$$ and because $\\sigma(x) = x + O(x^2) $ (meaning to be called \"<em>schlicht</em>\" ) also its formal power series inverse is of this form and we get\n$$ f°^h(1) = t + \\lambda^h c_\\lambda + O(\\lambda^{2h})$$ where $c_\\lambda = \\sigma(1/t-1) $<br>\nInterestingly, leaving $\\lambda$ indeterminate and $t=\\exp(\\lambda)$ unevaluated as formal power series Pari/GP gives $c_\\lambda$ as formal power series in the form as the OP in def. (6) provides its leading coefficients. </p>\n\n<hr>\n\n<blockquote>\n <p><strong>Derivation/implementation in Pari/GP</strong> </p>\n</blockquote>\n\n<p>Here I show, illustrated by sample calculations with Pari/GP, that there is no numerical approximation needed to get the coefficients at $\\lambda^k$ in definition (6). Instead there is a derivation on formal power series which provides exact rational coefficients. The method below reproduces perfectly your finding. </p>\n\n<p><hr>\n<strong>Preliminaries:</strong> a <a href=\"https://en.wikipedia.org/wiki/Carleman_matrix\" rel=\"nofollow noreferrer\">Carleman-matrix</a> <strong><em>B</em></strong> associated to a function $f(x)$ which has a power series with more than zero radius of convergence (and generalized: can be Euler-summed in a certain interval of its argument) is initially only a collection of column-vectors $B[,c]$ <em>($c$ indexing the c-olumn)</em>, which contain the coefficients of the formal power series of $f(x)$ and more precisely of $f(x)^c$ for $c=0 \\ldots \\infty$ in the column of index $c$ , and is thus in both direction of infinite size. It is an important special case when the power series of some function $f(x)$ has no constant term: then the shape of $B$ is lower triangular and the analytic handling of such Carlemanmatrices has been substantially established. We'll need this in the derivation below.<br>\n<sub> It is not trivial to show, that such a <em>collection</em> of columns-vectors can indeed be used as <strong><em>Matrix</em></strong>, applicable to concepts of matrix-algebra and -analysis, for instance raising to powers and especially diagonalization - but there is reliable literature about this so that we can omit such a step here. One pioneer is surely Eri Jabotinsky, see for instance this 1963-article <a href=\"http://www.jstor.org/stable/1993593\" rel=\"nofollow noreferrer\">\"Analytic iteration\"</a> (introducing the term \"representation matrix\")</sub> </p>\n\n<p>After the Carlemanmatrix <strong><em>B</em></strong> contains the coefficients of formal power series we need also a construct $V(x)$ as row vector $V(x)=[1,x,x^2,x^3,\\ldots]$ which I call of \"Vandermonde\"-type here. Using $V(x)$ as diagonal-matrix I pre-superscribe this like $ \\;^dV(x)$ \n<hr>\nFor instance the Carlemanmatrix $B_e$ for the function $f(x) = \\exp(x)$ looks like</p>\n\n<pre><code> 1 1 1 1 1 ...\n 0 1 2 3 4 ...\n 0 1/2! 2^2/2! 3^2/2! 4^2/2! ...\n 0 1/3! 2^3/3! 3^3/3! 4^3/3! ...\n 0 1/4! 2^4/4! 3^4/4! 4^4/4! ...\n ... ... ... ... .... ... \n</code></pre>\n\n<p>The clue with such Carlemanmatrices is, that the left-multiplication of some $B$ with a vector $V(x)$ as in\n $$V(x) \\cdot B = V(f(x)) \\tag 1$$ \ngives a resulting vector of the same structure $V(f(x)) = [ 1 , f(x), f(x)^2, f(x)^3, ... ]$ .<br>\nOf course, using the only \"relevant\" column in $B$ alone , then this gives the scalar expression $f(x)$: \n $$V(x) \\cdot B [,1] = f(x) \\tag {1.1}$$ </p>\n\n<p>This means: a composition of a function with itself $f(f(x))$ (in terms of its power series) is mapped to a <em>repeated multiplication</em> with its Carleman-matrix or in other words: is mapped to multiplication with <strong><em>powers</em></strong> of its Carlemanmatrix. </p>\n\n<p>The further clue is now, that for appropriate matrices a <strong><em>fractional</em></strong> power can be determined and for that cases we can thus easily define power series for the fractional iterates of the associated function. </p>\n\n<hr> \n\n<p><strong>Technically</strong> this means that the concept of diagonalization can be transferred to that Carleman-ansatz (of course for suitable Carlemanmatrices $B$). A case, for which this is widely accepted is the case where the Carleman-matrices for some function $f(x)$ is lower triangular.</p>\n\n<p>We have the function $f(x) = b^x $ and by the definition in the OP $1 \\lt b \\lt e^{1/e}$ . With this restriction, $f(x)$ has a fixpoint $t$ such that $f(t)=t$ (or: $b^t=t$ ) and also this fixpoint is attracting (which is important here). We can, as I wrote in the first part of text above, apply conjugacy to arrive at the function $g(x)=f(x`)´ = t^x-1 = ux + (ux)^2/2! + O(x^3) $ (for notational convenience I write $u$ instead of $\\lambda$ in the following). </p>\n\n<p>Given the lower triangular matrix of Stirlingnumbers $2$'nd kind $S2$ and its factorially similarity-scaled transform which I denote here as $fS2F$ we have the lower triangular Carlemanmatrix $U_t$ for $g(x)$ by $$ U_t = \\; ^dV(u) \\cdot fS2F \\tag {2.1}$$</p>\n\n<hr> \n\n<p>The top-left edge of $U_t$ looks like</p>\n\n<pre><code> 1 . . . . . \n 0 u/1! . . . . \n 0 u^2/2! u^2*2!/2! . . . \n 0 u^3/3! 3*u^3*2!/3! u^3*3!/3! . . \n 0 u^4/4! 7*u^4*2!/4! 6*u^4*3!/4! u^4*4!/4! . \n 0 u^5/5! 15*u^5*2!/5! 25*u^5*3!/5! 10*u^5*4!/5! u^5*5!/5! \n</code></pre>\n\n<p>or with cancelled entries </p>\n\n<pre><code> 1 . . . . .\n 0 u . . . .\n 0 1/2*u^2 u^2 . . .\n 0 1/6*u^3 u^3 u^3 . .\n 0 1/24*u^4 7/12*u^4 3/2*u^4 u^4 .\n 0 1/120*u^5 1/4*u^5 5/4*u^5 2*u^5 u^5\n</code></pre>\n\n<p>and taking the coefficients from the second column $$ g(x) = \\sum_{k=1}^\\infty u^k \\cdot \\frac{x^k}{ k!} \\tag {2.2}$$ \n<hr>\nThe formal power series for $h$'th fractional iterates $g°^h(x)$ are now generated by fractional powers of $U_t$ instead and this can be done by diagonalization: </p>\n\n<p>$$ U_t = M \\cdot D \\cdot W \\qquad \\qquad \\text{where I denote $W$ for } M^{-1} \\tag {3.1}$$\nwhere also the diagonal entries in $D$ are the consecutive powers of $u$\n$$ D = \\; ^d V(u) \\text{ and } D^h = \\; ^dV(u^h) \\tag {3.2}$$\nsuch that\n$$ U_t^h = M \\cdot \\; ^dV(u^h) \\cdot W \\tag {3.3} $$</p>\n\n<p>Unfortunately the current Pari/GP-version's diagonalization procedure cannot exploit the advantage of $U_t$ being triangular and also cannot keep the parameter $u$ symbolically. But it is not difficult to have a small recursive routine to generate $M$ , $D$ and $W (=M^{-1})$ for the triangular Carleman-case; the additional norming of $M$ to have a unit-diagonal is also included and with this the matrix $M$ is also of the Carleman-type (for details and an implementation in Pari/GP see the explanation in my linked article). The matrix <strong><em>M</em></strong> is now the Carlemanmatrix for the above introduced Schröder-function $\\sigma(x)$ which takes its coefficients from the second column in <strong><em>M</em></strong>. </p>\n\n<p>The top-left of the matrix $M$ looks like</p>\n\n<pre><code> 1 . . .\n 0 1 . .\n 0 u/(-2*u+2) 1 .\n 0 (2*u^3+u^2)/(6*u^3-6*u^2-6*u+6) u/(-u+1) 1\n</code></pre>\n\n<p>The entries in the (only relevant) $2$'nd column have a partially discernable pattern; I've described this in earlier discussions, see for instance <a href=\"http://go.helms-net.de/math/tetdocs/APT.htm\" rel=\"nofollow noreferrer\">here</a>, but which is too much to show at this place.<br>\n<hr>\nHaving this we can write for the $h$'th fractional iterate $g(x)$ of some $x$: </p>\n\n<p>$$ \\begin{array}{rll} V(x) \\cdot U_t^h &amp;= V( g°^h(x)) \\\\ &amp;= V(x) \\cdot (M \\cdot \\;^dV(u^h) \\cdot W) \\qquad \\qquad \\text{ and by associativity}\\\\\n &amp;= (V(x) \\cdot M) \\cdot \\;^dV(u^h) \\cdot W \\end{array} \\tag {4.1}$$</p>\n\n<p>Because $M$ is of Carlemantype we have also\n $$ V(x) \\cdot M = V(\\sigma(x)) \\tag {4.2}$$ \nor -using also conjugacy- $$ V(x´) \\cdot M = V(\\sigma(x´)) \\tag {4.3}$$\n and it is relevant here, that $\\sigma(x´)$ is <strong><em>identical to the (normed) Schröder-function</em></strong> for the iteration of $b^x$. </p>\n\n<p>In the problem as given in the OP we assume $x=1$ or $x´=1/t-1 = \\exp(-u)-1$ and to have the Schröder-function symbolically we have to look at\n$$ V(1´) \\cdot M = V(\\exp(-u)-1) \\cdot M = V(\\sigma(\\exp(-u)-1)) \\tag {4.4} $$ and this can still be done symbolically in $u$; having <strong><em>M</em></strong> in its symbolical representation, Pari/GP is able to give the leading part of $\\sigma(\\exp(-u)-1)$ by formal expansion of $\\exp(-u)$ as power series taken as argument of $\\sigma(\\cdot)$ with still rational coefficients. </p>\n\n<p>Having the left part of $(4.2)$ evaluated we need now\n$$ V(\\sigma(\\exp(-u)-1)) \\cdot \\; ^dV(u^h) \\cdot W[,1] = y =\\sigma°^{-1}(u^h \\cdot \\sigma(1´))\\tag {4.5}$$\nby the relevant $2$'nd column of $W$ only. </p>\n\n<p>Note that this gives a power series in $u^h$ with the (constant and rational) coefficients in $W[,1]$ and powers of the expression $\\sigma(\\exp(-u)-1)$ . To help Pari/GP to keep $u^h$ as a single symbol, we need to substitute it by the name of a unevaluated variable, say $u_h$ or $v$ and re-substitute it later after series-expansions.</p>\n\n<p>The final result $f°^h(1) $ occurs by the inverse conjugacy on $y$\n$$f°^h(1) = y` = (y+1)\\cdot t = y\\cdot t + t =t + y \\cdot \\exp(u) \\tag {4.6} $$\nwhich can still be expressed as formal power series in the indeterminate $u$ and $u^h$. </p>\n\n<p>The OP asks now for the first coefficient in $y \\cdot \\exp(u) $ and namely for its series expression when $u$ <em>and</em> $v (=u^h) $ are both kept as indeterminate parameter. Pari/GP is still able to evaluate the last expression symbolically and arrives at the same set of coefficients as given in eq (6) in the OP.\n<hr>\n<strong>Additional computation:</strong> a short recalculation gives also the leading terms of the powerseries of the following coefficients in (5) of the OP, which I index here as $c_{\\lambda,k}$ : </p>\n\n<p>$$\\small \\begin{array} {r}\nc_{\\lambda,1}=&amp; - 1 u^1/1!&amp; - 1 u^2/2!&amp; + 2 u^3/3!&amp; + 11 u^4/4!&amp; + 44 u^5/5!&amp; + 89 u^6/6!&amp; + 636 u^7/7!&amp; + O(u^8) \\\\\nc_{\\lambda,2}=&amp;&amp;&amp; - 3 u^3/3!&amp; - 12 u^4/4!&amp; - 5 u^5/5!&amp; + 150 u^6/6!&amp; + 1463 u^7/7!&amp; + O(u^8) \\\\\nc_{\\lambda,3}=&amp;&amp;&amp;&amp;&amp; - 40 u^5/5!&amp; - 240 u^6/6!&amp; - 840 u^7/7!&amp; + O(u^8) \\\\\nc_{\\lambda,4}=&amp;&amp;&amp;&amp;&amp;&amp;&amp; - 1260 u^7/7!&amp; + O(u^8) \\\\\n \\end{array}\n$$\nand recalling that $\\; ^hb $ is my notation for your $ \\; ^na $ we get\n$$ \\; ^na = \\; ^hb = t +c_{\\lambda,1} u^h + c_{\\lambda,2} u^{2h} + c_{\\lambda,3} u^{3h} + \\cdots \\tag 5$$\nRemark: I don't think that a closing term $O(u^{4h})$ makes really sense here because we assume $h&gt;-2$ to be of <em>any real value</em>, especially fractional, and I'm not firm with using that notation in such cases.</p>\n" }, { "answer_id": 272150, "score": 5, "text": "<p>Let $$c_n(\\lambda)=\\frac{e^\\lambda-{^n a}}{ \\lambda^n},\\quad d_n(\\lambda)=e^{-\\lambda}c_n(\\lambda)$$ ($d_n$ seems to be simpler). In particular initial functions $c_0(\\lambda)=e^\\lambda-1$ and $d_0(\\lambda)=1-e^{-\\lambda}\\sim\\lambda$ (as $\\lambda\\to 0$) have good power series of the variable $\\lambda$. We are intersted in the limits $$c_\\lambda=\\lim\\limits_{n\\to\\infty}c_n(\\lambda),\\quad d_\\lambda=\\lim\\limits_{n\\to\\infty}d_n(\\lambda).$$ The recurrence\n\\begin{gather*}\nd_{n+1}(\\lambda)=\\frac{1-e^{-\\lambda^{n+1}d_n(\\lambda)}}{\\lambda^{n+1} }\n\\end{gather*}\nfollows from the definition. By induction $d_n(\\lambda)\\sim\\lambda$. Expanding the recurrence above we get\n\\begin{gather*}\nd_{n+1}(\\lambda)=d_n(\\lambda)-\\frac{\\lambda^{n+1}d_n^2(\\lambda)}{2}+\\ldots\\equiv d_n(\\lambda)\\pmod{\\lambda^{n+3} }.\n\\end{gather*}\nIt means that the limit $d_\\lambda$ (as a power series) does exist.</p>\n\n<p><strong>EDT.</strong> Numerators $a(k)$ of $c_\\lambda$-coefficients are not very random. At least first $41$ coefficients from OP satisfy the congruences\n\\begin{gather*}\na(k)\\equiv\\begin{cases}\n0,&amp;\\text{if } k\\equiv 1\\pmod{2 },\\\\\n1,&amp;\\text{if }k\\equiv 0\\pmod{4 },\\\\\n3,&amp;\\text{if }k\\equiv 2\\pmod{4 },\\\\\n\\end{cases}\\pmod{4 }\\quad(k\\ge 4); \\\\\na(k)\\equiv\\begin{cases}\n0,&amp;\\text{if }k\\equiv 1,2\\pmod{6 },\\\\\n1,&amp;\\text{if }k\\equiv0,3,4,5\\pmod{6 },\\\\\n\\end{cases}\\pmod{3 }\\quad(k\\ge 3).\n\\end{gather*}</p>\n" }, { "answer_id": 272338, "score": 2, "text": "<p><strong><em>This is not an answer, only an extended comment at V. Reshetnikov's request in his comment <a href=\"https://mathoverflow.net/a/272157/7710\">at my answer</a></em></strong> </p>\n\n<p>@VladimirReshetnikov - not so far now. I think this problem of a closed form\nis also intimately related to the so far unsolved question of the \"mag\"-numbers as discussed in \n<a href=\"https://mathoverflow.net/questions/57627\">\"Is there an “elegant” non-recursive formula for these coefficients?\"</a> \nwhich is focusing the series expansion for the Schröder-function $\\sigma()$ as I've used it here.</p>\n\n<p>Also your approach using q-binomials for a series <a href=\"https://mathoverflow.net/questions/259278\">\"An explicit series representation for the analytic tetration with complex height\"</a> was/is much interesting for me - but although I had similar analyses involving q-binomials I didn't arrive at any conciser formulae than the formal algebraic descriptions following from the matrix-based concepts as discussed, for instance, here: <a href=\"http://go.helms-net.de/math/tetdocs/ExponentialPolynomialInterpolation.pdf\" rel=\"nofollow noreferrer\">\"Exponential polynomial interpolation\"</a> and also here <a href=\"http://go.helms-net.de/math/tetdocs/BinomialDiagonalization.htm\" rel=\"nofollow noreferrer\">\"Relation between binomial- and diagonalization method\"</a> but at least showing at its end the series <strong><em>(5)</em></strong> in your OP assuming $\\lambda=\\log 2$ with evaluated coefficients in real numbers and which gives also a short resume of the -as I christened it- \"exponential polynomial interpolation\" method which involves the q-binomials.</p>\n\n<p>In general: the -in my opinion- most sophisticated explication of the coefficients of the function $t^x-1$ (and by conjugacy as well of $b^x$ ) in terms of the logarithm $u$ of $t$ and of $u^h$ comes via the diagonalization-approach as shown in <a href=\"http://go.helms-net.de/math/tetdocs/APT.htm\" rel=\"nofollow noreferrer\">\"Eigendecomposition of triangular matrix-operators (here: Ut)\"</a> from where the coefficients of your formulae <strong><em>(5)</em></strong> and <strong><em>(6)</em></strong> can be derived, and even symbolically<br>\nBut as well as with the q-binomial approach I did not yet find any simplifications/closed forms besides that (recursive)\nbasic generating schemes for that formulae/coefficients, I am really sorry. </p>\n\n<p>Of course, in case I'll find something worth sometime I'll surely link to that discussion here.</p>\n" } ]
Question: A curious series related to the asymptotic behavior of the tetration Answer 1: <p>The coefficients of the series (6) can be found symbolically in $\lambda$ and rational numbers. </p> <p>Consider some base $b$ with $1 \le b \le e^{1/e}$ , the fixpoint $t$ for $f(x)=b^x $ and $f(t)=t$. For bases $b$ in this interval we have a real fixpoint $t$ with $1 \le t \le e$ and $\lambda = \log(t) $ with $0 \le \lambda \le 1$ with $b^t = t$ <em>(which is also attracting when $\lambda,t,b$ are not on the bounds of the resp. intervals)</em>. </p> <p>$\ $ </p> <p>We assume now a conjugate function $d(z)=z/t-1$ and $d°^{-1}(z)=(z+1)\cdot t$ which I write here for shortness in sup-suffix notation $z´ = d(z) $ and $z`=d°^{-1}(z)$ .</p> <p>The <strong><em><a href="https://en.wikipedia.org/wiki/Schr%C3%B6der%27s_equation" rel="nofollow noreferrer">Schröder-mechanism</a></em></strong>, using this conjugacy, allows an invertible "<em>schlicht</em>" function $\sigma(x)$ such that $$ f°^h(x) = \left(\sigma°^{-1}( \lambda^h \cdot \sigma(x´))\right)`$$ </p> <p>Now the given problem assumes $x=1$ such that $x´ = 1/t-1 = \exp(-\lambda)-1$ and we have explicitely $$ f°^h(1) = \left(\sigma°^{-1}( \lambda^h \cdot \sigma(\exp(-\lambda)-1))\right) \cdot \exp(\lambda) + t$$ and because $\sigma(x) = x + O(x^2) $ (meaning to be called "<em>schlicht</em>" ) also its formal power series inverse is of this form and we get $$ f°^h(1) = t + \lambda^h c_\lambda + O(\lambda^{2h})$$ where $c_\lambda = \sigma(1/t-1) $<br> Interestingly, leaving $\lambda$ indeterminate and $t=\exp(\lambda)$ unevaluated as formal power series Pari/GP gives $c_\lambda$ as formal power series in the form as the OP in def. (6) provides its leading coefficients. </p> <hr> <blockquote> <p><strong>Derivation/implementation in Pari/GP</strong> </p> </blockquote> <p>Here I show, illustrated by sample calculations with Pari/GP, that there is no numerical approximation needed to get the coefficients at $\lambda^k$ in definition (6). Instead there is a derivation on formal power series which provides exact rational coefficients. The method below reproduces perfectly your finding. </p> <p><hr> <strong>Preliminaries:</strong> a <a href="https://en.wikipedia.org/wiki/Carleman_matrix" rel="nofollow noreferrer">Carleman-matrix</a> <strong><em>B</em></strong> associated to a function $f(x)$ which has a power series with more than zero radius of convergence (and generalized: can be Euler-summed in a certain interval of its argument) is initially only a collection of column-vectors $B[,c]$ <em>($c$ indexing the c-olumn)</em>, which contain the coefficients of the formal power series of $f(x)$ and more precisely of $f(x)^c$ for $c=0 \ldots \infty$ in the column of index $c$ , and is thus in both direction of infinite size. It is an important special case when the power series of some function $f(x)$ has no constant term: then the shape of $B$ is lower triangular and the analytic handling of such Carlemanmatrices has been substantially established. We'll need this in the derivation below.<br> <sub> It is not trivial to show, that such a <em>collection</em> of columns-vectors can indeed be used as <strong><em>Matrix</em></strong>, applicable to concepts of matrix-algebra and -analysis, for instance raising to powers and especially diagonalization - but there is reliable literature about this so that we can omit such a step here. One pioneer is surely Eri Jabotinsky, see for instance this 1963-article <a href="http://www.jstor.org/stable/1993593" rel="nofollow noreferrer">"Analytic iteration"</a> (introducing the term "representation matrix")</sub> </p> <p>After the Carlemanmatrix <strong><em>B</em></strong> contains the coefficients of formal power series we need also a construct $V(x)$ as row vector $V(x)=[1,x,x^2,x^3,\ldots]$ which I call of "Vandermonde"-type here. Using $V(x)$ as diagonal-matrix I pre-superscribe this like $ \;^dV(x)$ <hr> For instance the Carlemanmatrix $B_e$ for the function $f(x) = \exp(x)$ looks like</p> <pre><code> 1 1 1 1 1 ... 0 1 2 3 4 ... 0 1/2! 2^2/2! 3^2/2! 4^2/2! ... 0 1/3! 2^3/3! 3^3/3! 4^3/3! ... 0 1/4! 2^4/4! 3^4/4! 4^4/4! ... ... ... ... ... .... ... </code></pre> <p>The clue with such Carlemanmatrices is, that the left-multiplication of some $B$ with a vector $V(x)$ as in $$V(x) \cdot B = V(f(x)) \tag 1$$ gives a resulting vector of the same structure $V(f(x)) = [ 1 , f(x), f(x)^2, f(x)^3, ... ]$ .<br> Of course, using the only "relevant" column in $B$ alone , then this gives the scalar expression $f(x)$: $$V(x) \cdot B [,1] = f(x) \tag {1.1}$$ </p> <p>This means: a composition of a function with itself $f(f(x))$ (in terms of its power series) is mapped to a <em>repeated multiplication</em> with its Carleman-matrix or in other words: is mapped to multiplication with <strong><em>powers</em></strong> of its Carlemanmatrix. </p> <p>The further clue is now, that for appropriate matrices a <strong><em>fractional</em></strong> power can be determined and for that cases we can thus easily define power series for the fractional iterates of the associated function. </p> <hr> <p><strong>Technically</strong> this means that the concept of diagonalization can be transferred to that Carleman-ansatz (of course for suitable Carlemanmatrices $B$). A case, for which this is widely accepted is the case where the Carleman-matrices for some function $f(x)$ is lower triangular.</p> <p>We have the function $f(x) = b^x $ and by the definition in the OP $1 \lt b \lt e^{1/e}$ . With this restriction, $f(x)$ has a fixpoint $t$ such that $f(t)=t$ (or: $b^t=t$ ) and also this fixpoint is attracting (which is important here). We can, as I wrote in the first part of text above, apply conjugacy to arrive at the function $g(x)=f(x`)´ = t^x-1 = ux + (ux)^2/2! + O(x^3) $ (for notational convenience I write $u$ instead of $\lambda$ in the following). </p> <p>Given the lower triangular matrix of Stirlingnumbers $2$'nd kind $S2$ and its factorially similarity-scaled transform which I denote here as $fS2F$ we have the lower triangular Carlemanmatrix $U_t$ for $g(x)$ by $$ U_t = \; ^dV(u) \cdot fS2F \tag {2.1}$$</p> <hr> <p>The top-left edge of $U_t$ looks like</p> <pre><code> 1 . . . . . 0 u/1! . . . . 0 u^2/2! u^2*2!/2! . . . 0 u^3/3! 3*u^3*2!/3! u^3*3!/3! . . 0 u^4/4! 7*u^4*2!/4! 6*u^4*3!/4! u^4*4!/4! . 0 u^5/5! 15*u^5*2!/5! 25*u^5*3!/5! 10*u^5*4!/5! u^5*5!/5! </code></pre> <p>or with cancelled entries </p> <pre><code> 1 . . . . . 0 u . . . . 0 1/2*u^2 u^2 . . . 0 1/6*u^3 u^3 u^3 . . 0 1/24*u^4 7/12*u^4 3/2*u^4 u^4 . 0 1/120*u^5 1/4*u^5 5/4*u^5 2*u^5 u^5 </code></pre> <p>and taking the coefficients from the second column $$ g(x) = \sum_{k=1}^\infty u^k \cdot \frac{x^k}{ k!} \tag {2.2}$$ <hr> The formal power series for $h$'th fractional iterates $g°^h(x)$ are now generated by fractional powers of $U_t$ instead and this can be done by diagonalization: </p> <p>$$ U_t = M \cdot D \cdot W \qquad \qquad \text{where I denote $W$ for } M^{-1} \tag {3.1}$$ where also the diagonal entries in $D$ are the consecutive powers of $u$ $$ D = \; ^d V(u) \text{ and } D^h = \; ^dV(u^h) \tag {3.2}$$ such that $$ U_t^h = M \cdot \; ^dV(u^h) \cdot W \tag {3.3} $$</p> <p>Unfortunately the current Pari/GP-version's diagonalization procedure cannot exploit the advantage of $U_t$ being triangular and also cannot keep the parameter $u$ symbolically. But it is not difficult to have a small recursive routine to generate $M$ , $D$ and $W (=M^{-1})$ for the triangular Carleman-case; the additional norming of $M$ to have a unit-diagonal is also included and with this the matrix $M$ is also of the Carleman-type (for details and an implementation in Pari/GP see the explanation in my linked article). The matrix <strong><em>M</em></strong> is now the Carlemanmatrix for the above introduced Schröder-function $\sigma(x)$ which takes its coefficients from the second column in <strong><em>M</em></strong>. </p> <p>The top-left of the matrix $M$ looks like</p> <pre><code> 1 . . . 0 1 . . 0 u/(-2*u+2) 1 . 0 (2*u^3+u^2)/(6*u^3-6*u^2-6*u+6) u/(-u+1) 1 </code></pre> <p>The entries in the (only relevant) $2$'nd column have a partially discernable pattern; I've described this in earlier discussions, see for instance <a href="http://go.helms-net.de/math/tetdocs/APT.htm" rel="nofollow noreferrer">here</a>, but which is too much to show at this place.<br> <hr> Having this we can write for the $h$'th fractional iterate $g(x)$ of some $x$: </p> <p>$$ \begin{array}{rll} V(x) \cdot U_t^h &amp;= V( g°^h(x)) \\ &amp;= V(x) \cdot (M \cdot \;^dV(u^h) \cdot W) \qquad \qquad \text{ and by associativity}\\ &amp;= (V(x) \cdot M) \cdot \;^dV(u^h) \cdot W \end{array} \tag {4.1}$$</p> <p>Because $M$ is of Carlemantype we have also $$ V(x) \cdot M = V(\sigma(x)) \tag {4.2}$$ or -using also conjugacy- $$ V(x´) \cdot M = V(\sigma(x´)) \tag {4.3}$$ and it is relevant here, that $\sigma(x´)$ is <strong><em>identical to the (normed) Schröder-function</em></strong> for the iteration of $b^x$. </p> <p>In the problem as given in the OP we assume $x=1$ or $x´=1/t-1 = \exp(-u)-1$ and to have the Schröder-function symbolically we have to look at $$ V(1´) \cdot M = V(\exp(-u)-1) \cdot M = V(\sigma(\exp(-u)-1)) \tag {4.4} $$ and this can still be done symbolically in $u$; having <strong><em>M</em></strong> in its symbolical representation, Pari/GP is able to give the leading part of $\sigma(\exp(-u)-1)$ by formal expansion of $\exp(-u)$ as power series taken as argument of $\sigma(\cdot)$ with still rational coefficients. </p> <p>Having the left part of $(4.2)$ evaluated we need now $$ V(\sigma(\exp(-u)-1)) \cdot \; ^dV(u^h) \cdot W[,1] = y =\sigma°^{-1}(u^h \cdot \sigma(1´))\tag {4.5}$$ by the relevant $2$'nd column of $W$ only. </p> <p>Note that this gives a power series in $u^h$ with the (constant and rational) coefficients in $W[,1]$ and powers of the expression $\sigma(\exp(-u)-1)$ . To help Pari/GP to keep $u^h$ as a single symbol, we need to substitute it by the name of a unevaluated variable, say $u_h$ or $v$ and re-substitute it later after series-expansions.</p> <p>The final result $f°^h(1) $ occurs by the inverse conjugacy on $y$ $$f°^h(1) = y` = (y+1)\cdot t = y\cdot t + t =t + y \cdot \exp(u) \tag {4.6} $$ which can still be expressed as formal power series in the indeterminate $u$ and $u^h$. </p> <p>The OP asks now for the first coefficient in $y \cdot \exp(u) $ and namely for its series expression when $u$ <em>and</em> $v (=u^h) $ are both kept as indeterminate parameter. Pari/GP is still able to evaluate the last expression symbolically and arrives at the same set of coefficients as given in eq (6) in the OP. <hr> <strong>Additional computation:</strong> a short recalculation gives also the leading terms of the powerseries of the following coefficients in (5) of the OP, which I index here as $c_{\lambda,k}$ : </p> <p>$$\small \begin{array} {r} c_{\lambda,1}=&amp; - 1 u^1/1!&amp; - 1 u^2/2!&amp; + 2 u^3/3!&amp; + 11 u^4/4!&amp; + 44 u^5/5!&amp; + 89 u^6/6!&amp; + 636 u^7/7!&amp; + O(u^8) \\ c_{\lambda,2}=&amp;&amp;&amp; - 3 u^3/3!&amp; - 12 u^4/4!&amp; - 5 u^5/5!&amp; + 150 u^6/6!&amp; + 1463 u^7/7!&amp; + O(u^8) \\ c_{\lambda,3}=&amp;&amp;&amp;&amp;&amp; - 40 u^5/5!&amp; - 240 u^6/6!&amp; - 840 u^7/7!&amp; + O(u^8) \\ c_{\lambda,4}=&amp;&amp;&amp;&amp;&amp;&amp;&amp; - 1260 u^7/7!&amp; + O(u^8) \\ \end{array} $$ and recalling that $\; ^hb $ is my notation for your $ \; ^na $ we get $$ \; ^na = \; ^hb = t +c_{\lambda,1} u^h + c_{\lambda,2} u^{2h} + c_{\lambda,3} u^{3h} + \cdots \tag 5$$ Remark: I don't think that a closing term $O(u^{4h})$ makes really sense here because we assume $h&gt;-2$ to be of <em>any real value</em>, especially fractional, and I'm not firm with using that notation in such cases.</p> Answer 2: <p>Let $$c_n(\lambda)=\frac{e^\lambda-{^n a}}{ \lambda^n},\quad d_n(\lambda)=e^{-\lambda}c_n(\lambda)$$ ($d_n$ seems to be simpler). In particular initial functions $c_0(\lambda)=e^\lambda-1$ and $d_0(\lambda)=1-e^{-\lambda}\sim\lambda$ (as $\lambda\to 0$) have good power series of the variable $\lambda$. We are intersted in the limits $$c_\lambda=\lim\limits_{n\to\infty}c_n(\lambda),\quad d_\lambda=\lim\limits_{n\to\infty}d_n(\lambda).$$ The recurrence \begin{gather*} d_{n+1}(\lambda)=\frac{1-e^{-\lambda^{n+1}d_n(\lambda)}}{\lambda^{n+1} } \end{gather*} follows from the definition. By induction $d_n(\lambda)\sim\lambda$. Expanding the recurrence above we get \begin{gather*} d_{n+1}(\lambda)=d_n(\lambda)-\frac{\lambda^{n+1}d_n^2(\lambda)}{2}+\ldots\equiv d_n(\lambda)\pmod{\lambda^{n+3} }. \end{gather*} It means that the limit $d_\lambda$ (as a power series) does exist.</p> <p><strong>EDT.</strong> Numerators $a(k)$ of $c_\lambda$-coefficients are not very random. At least first $41$ coefficients from OP satisfy the congruences \begin{gather*} a(k)\equiv\begin{cases} 0,&amp;\text{if } k\equiv 1\pmod{2 },\\ 1,&amp;\text{if }k\equiv 0\pmod{4 },\\ 3,&amp;\text{if }k\equiv 2\pmod{4 },\\ \end{cases}\pmod{4 }\quad(k\ge 4); \\ a(k)\equiv\begin{cases} 0,&amp;\text{if }k\equiv 1,2\pmod{6 },\\ 1,&amp;\text{if }k\equiv0,3,4,5\pmod{6 },\\ \end{cases}\pmod{3 }\quad(k\ge 3). \end{gather*}</p> Answer 3: <p><strong><em>This is not an answer, only an extended comment at V. Reshetnikov's request in his comment <a href="https://mathoverflow.net/a/272157/7710">at my answer</a></em></strong> </p> <p>@VladimirReshetnikov - not so far now. I think this problem of a closed form is also intimately related to the so far unsolved question of the "mag"-numbers as discussed in <a href="https://mathoverflow.net/questions/57627">"Is there an “elegant” non-recursive formula for these coefficients?"</a> which is focusing the series expansion for the Schröder-function $\sigma()$ as I've used it here.</p> <p>Also your approach using q-binomials for a series <a href="https://mathoverflow.net/questions/259278">"An explicit series representation for the analytic tetration with complex height"</a> was/is much interesting for me - but although I had similar analyses involving q-binomials I didn't arrive at any conciser formulae than the formal algebraic descriptions following from the matrix-based concepts as discussed, for instance, here: <a href="http://go.helms-net.de/math/tetdocs/ExponentialPolynomialInterpolation.pdf" rel="nofollow noreferrer">"Exponential polynomial interpolation"</a> and also here <a href="http://go.helms-net.de/math/tetdocs/BinomialDiagonalization.htm" rel="nofollow noreferrer">"Relation between binomial- and diagonalization method"</a> but at least showing at its end the series <strong><em>(5)</em></strong> in your OP assuming $\lambda=\log 2$ with evaluated coefficients in real numbers and which gives also a short resume of the -as I christened it- "exponential polynomial interpolation" method which involves the q-binomials.</p> <p>In general: the -in my opinion- most sophisticated explication of the coefficients of the function $t^x-1$ (and by conjugacy as well of $b^x$ ) in terms of the logarithm $u$ of $t$ and of $u^h$ comes via the diagonalization-approach as shown in <a href="http://go.helms-net.de/math/tetdocs/APT.htm" rel="nofollow noreferrer">"Eigendecomposition of triangular matrix-operators (here: Ut)"</a> from where the coefficients of your formulae <strong><em>(5)</em></strong> and <strong><em>(6)</em></strong> can be derived, and even symbolically<br> But as well as with the q-binomial approach I did not yet find any simplifications/closed forms besides that (recursive) basic generating schemes for that formulae/coefficients, I am really sorry. </p> <p>Of course, in case I'll find something worth sometime I'll surely link to that discussion here.</p> Answer 4: <p>The coefficients of the series (6) can be found symbolically in $\lambda$ and rational numbers. </p> <p>Consider some base $b$ with $1 \le b \le e^{1/e}$ , the fixpoint $t$ for $f(x)=b^x $ and $f(t)=t$. For bases $b$ in this interval we have a real fixpoint $t$ with $1 \le t \le e$ and $\lambda = \log(t) $ with $0 \le \lambda \le 1$ with $b^t = t$ <em>(which is also attracting when $\lambda,t,b$ are not on the bounds of the resp. intervals)</em>. </p> <p>$\ $ </p> <p>We assume now a conjugate function $d(z)=z/t-1$ and $d°^{-1}(z)=(z+1)\cdot t$ which I write here for shortness in sup-suffix notation $z´ = d(z) $ and $z`=d°^{-1}(z)$ .</p> <p>The <strong><em><a href="https://en.wikipedia.org/wiki/Schr%C3%B6der%27s_equation" rel="nofollow noreferrer">Schröder-mechanism</a></em></strong>, using this conjugacy, allows an invertible "<em>schlicht</em>" function $\sigma(x)$ such that $$ f°^h(x) = \left(\sigma°^{-1}( \lambda^h \cdot \sigma(x´))\right)`$$ </p> <p>Now the given problem assumes $x=1$ such that $x´ = 1/t-1 = \exp(-\lambda)-1$ and we have explicitely $$ f°^h(1) = \left(\sigma°^{-1}( \lambda^h \cdot \sigma(\exp(-\lambda)-1))\right) \cdot \exp(\lambda) + t$$ and because $\sigma(x) = x + O(x^2) $ (meaning to be called "<em>schlicht</em>" ) also its formal power series inverse is of this form and we get $$ f°^h(1) = t + \lambda^h c_\lambda + O(\lambda^{2h})$$ where $c_\lambda = \sigma(1/t-1) $<br> Interestingly, leaving $\lambda$ indeterminate and $t=\exp(\lambda)$ unevaluated as formal power series Pari/GP gives $c_\lambda$ as formal power series in the form as the OP in def. (6) provides its leading coefficients. </p> <hr> <blockquote> <p><strong>Derivation/implementation in Pari/GP</strong> </p> </blockquote> <p>Here I show, illustrated by sample calculations with Pari/GP, that there is no numerical approximation needed to get the coefficients at $\lambda^k$ in definition (6). Instead there is a derivation on formal power series which provides exact rational coefficients. The method below reproduces perfectly your finding. </p> <p><hr> <strong>Preliminaries:</strong> a <a href="https://en.wikipedia.org/wiki/Carleman_matrix" rel="nofollow noreferrer">Carleman-matrix</a> <strong><em>B</em></strong> associated to a function $f(x)$ which has a power series with more than zero radius of convergence (and generalized: can be Euler-summed in a certain interval of its argument) is initially only a collection of column-vectors $B[,c]$ <em>($c$ indexing the c-olumn)</em>, which contain the coefficients of the formal power series of $f(x)$ and more precisely of $f(x)^c$ for $c=0 \ldots \infty$ in the column of index $c$ , and is thus in both direction of infinite size. It is an important special case when the power series of some function $f(x)$ has no constant term: then the shape of $B$ is lower triangular and the analytic handling of such Carlemanmatrices has been substantially established. We'll need this in the derivation below.<br> <sub> It is not trivial to show, that such a <em>collection</em> of columns-vectors can indeed be used as <strong><em>Matrix</em></strong>, applicable to concepts of matrix-algebra and -analysis, for instance raising to powers and especially diagonalization - but there is reliable literature about this so that we can omit such a step here. One pioneer is surely Eri Jabotinsky, see for instance this 1963-article <a href="http://www.jstor.org/stable/1993593" rel="nofollow noreferrer">"Analytic iteration"</a> (introducing the term "representation matrix")</sub> </p> <p>After the Carlemanmatrix <strong><em>B</em></strong> contains the coefficients of formal power series we need also a construct $V(x)$ as row vector $V(x)=[1,x,x^2,x^3,\ldots]$ which I call of "Vandermonde"-type here. Using $V(x)$ as diagonal-matrix I pre-superscribe this like $ \;^dV(x)$ <hr> For instance the Carlemanmatrix $B_e$ for the function $f(x) = \exp(x)$ looks like</p> <pre><code> 1 1 1 1 1 ... 0 1 2 3 4 ... 0 1/2! 2^2/2! 3^2/2! 4^2/2! ... 0 1/3! 2^3/3! 3^3/3! 4^3/3! ... 0 1/4! 2^4/4! 3^4/4! 4^4/4! ... ... ... ... ... .... ... </code></pre> <p>The clue with such Carlemanmatrices is, that the left-multiplication of some $B$ with a vector $V(x)$ as in $$V(x) \cdot B = V(f(x)) \tag 1$$ gives a resulting vector of the same structure $V(f(x)) = [ 1 , f(x), f(x)^2, f(x)^3, ... ]$ .<br> Of course, using the only "relevant" column in $B$ alone , then this gives the scalar expression $f(x)$: $$V(x) \cdot B [,1] = f(x) \tag {1.1}$$ </p> <p>This means: a composition of a function with itself $f(f(x))$ (in terms of its power series) is mapped to a <em>repeated multiplication</em> with its Carleman-matrix or in other words: is mapped to multiplication with <strong><em>powers</em></strong> of its Carlemanmatrix. </p> <p>The further clue is now, that for appropriate matrices a <strong><em>fractional</em></strong> power can be determined and for that cases we can thus easily define power series for the fractional iterates of the associated function. </p> <hr> <p><strong>Technically</strong> this means that the concept of diagonalization can be transferred to that Carleman-ansatz (of course for suitable Carlemanmatrices $B$). A case, for which this is widely accepted is the case where the Carleman-matrices for some function $f(x)$ is lower triangular.</p> <p>We have the function $f(x) = b^x $ and by the definition in the OP $1 \lt b \lt e^{1/e}$ . With this restriction, $f(x)$ has a fixpoint $t$ such that $f(t)=t$ (or: $b^t=t$ ) and also this fixpoint is attracting (which is important here). We can, as I wrote in the first part of text above, apply conjugacy to arrive at the function $g(x)=f(x`)´ = t^x-1 = ux + (ux)^2/2! + O(x^3) $ (for notational convenience I write $u$ instead of $\lambda$ in the following). </p> <p>Given the lower triangular matrix of Stirlingnumbers $2$'nd kind $S2$ and its factorially similarity-scaled transform which I denote here as $fS2F$ we have the lower triangular Carlemanmatrix $U_t$ for $g(x)$ by $$ U_t = \; ^dV(u) \cdot fS2F \tag {2.1}$$</p> <hr> <p>The top-left edge of $U_t$ looks like</p> <pre><code> 1 . . . . . 0 u/1! . . . . 0 u^2/2! u^2*2!/2! . . . 0 u^3/3! 3*u^3*2!/3! u^3*3!/3! . . 0 u^4/4! 7*u^4*2!/4! 6*u^4*3!/4! u^4*4!/4! . 0 u^5/5! 15*u^5*2!/5! 25*u^5*3!/5! 10*u^5*4!/5! u^5*5!/5! </code></pre> <p>or with cancelled entries </p> <pre><code> 1 . . . . . 0 u . . . . 0 1/2*u^2 u^2 . . . 0 1/6*u^3 u^3 u^3 . . 0 1/24*u^4 7/12*u^4 3/2*u^4 u^4 . 0 1/120*u^5 1/4*u^5 5/4*u^5 2*u^5 u^5 </code></pre> <p>and taking the coefficients from the second column $$ g(x) = \sum_{k=1}^\infty u^k \cdot \frac{x^k}{ k!} \tag {2.2}$$ <hr> The formal power series for $h$'th fractional iterates $g°^h(x)$ are now generated by fractional powers of $U_t$ instead and this can be done by diagonalization: </p> <p>$$ U_t = M \cdot D \cdot W \qquad \qquad \text{where I denote $W$ for } M^{-1} \tag {3.1}$$ where also the diagonal entries in $D$ are the consecutive powers of $u$ $$ D = \; ^d V(u) \text{ and } D^h = \; ^dV(u^h) \tag {3.2}$$ such that $$ U_t^h = M \cdot \; ^dV(u^h) \cdot W \tag {3.3} $$</p> <p>Unfortunately the current Pari/GP-version's diagonalization procedure cannot exploit the advantage of $U_t$ being triangular and also cannot keep the parameter $u$ symbolically. But it is not difficult to have a small recursive routine to generate $M$ , $D$ and $W (=M^{-1})$ for the triangular Carleman-case; the additional norming of $M$ to have a unit-diagonal is also included and with this the matrix $M$ is also of the Carleman-type (for details and an implementation in Pari/GP see the explanation in my linked article). The matrix <strong><em>M</em></strong> is now the Carlemanmatrix for the above introduced Schröder-function $\sigma(x)$ which takes its coefficients from the second column in <strong><em>M</em></strong>. </p> <p>The top-left of the matrix $M$ looks like</p> <pre><code> 1 . . . 0 1 . . 0 u/(-2*u+2) 1 . 0 (2*u^3+u^2)/(6*u^3-6*u^2-6*u+6) u/(-u+1) 1 </code></pre> <p>The entries in the (only relevant) $2$'nd column have a partially discernable pattern; I've described this in earlier discussions, see for instance <a href="http://go.helms-net.de/math/tetdocs/APT.htm" rel="nofollow noreferrer">here</a>, but which is too much to show at this place.<br> <hr> Having this we can write for the $h$'th fractional iterate $g(x)$ of some $x$: </p> <p>$$ \begin{array}{rll} V(x) \cdot U_t^h &amp;= V( g°^h(x)) \\ &amp;= V(x) \cdot (M \cdot \;^dV(u^h) \cdot W) \qquad \qquad \text{ and by associativity}\\ &amp;= (V(x) \cdot M) \cdot \;^dV(u^h) \cdot W \end{array} \tag {4.1}$$</p> <p>Because $M$ is of Carlemantype we have also $$ V(x) \cdot M = V(\sigma(x)) \tag {4.2}$$ or -using also conjugacy- $$ V(x´) \cdot M = V(\sigma(x´)) \tag {4.3}$$ and it is relevant here, that $\sigma(x´)$ is <strong><em>identical to the (normed) Schröder-function</em></strong> for the iteration of $b^x$. </p> <p>In the problem as given in the OP we assume $x=1$ or $x´=1/t-1 = \exp(-u)-1$ and to have the Schröder-function symbolically we have to look at $$ V(1´) \cdot M = V(\exp(-u)-1) \cdot M = V(\sigma(\exp(-u)-1)) \tag {4.4} $$ and this can still be done symbolically in $u$; having <strong><em>M</em></strong> in its symbolical representation, Pari/GP is able to give the leading part of $\sigma(\exp(-u)-1)$ by formal expansion of $\exp(-u)$ as power series taken as argument of $\sigma(\cdot)$ with still rational coefficients. </p> <p>Having the left part of $(4.2)$ evaluated we need now $$ V(\sigma(\exp(-u)-1)) \cdot \; ^dV(u^h) \cdot W[,1] = y =\sigma°^{-1}(u^h \cdot \sigma(1´))\tag {4.5}$$ by the relevant $2$'nd column of $W$ only. </p> <p>Note that this gives a power series in $u^h$ with the (constant and rational) coefficients in $W[,1]$ and powers of the expression $\sigma(\exp(-u)-1)$ . To help Pari/GP to keep $u^h$ as a single symbol, we need to substitute it by the name of a unevaluated variable, say $u_h$ or $v$ and re-substitute it later after series-expansions.</p> <p>The final result $f°^h(1) $ occurs by the inverse conjugacy on $y$ $$f°^h(1) = y` = (y+1)\cdot t = y\cdot t + t =t + y \cdot \exp(u) \tag {4.6} $$ which can still be expressed as formal power series in the indeterminate $u$ and $u^h$. </p> <p>The OP asks now for the first coefficient in $y \cdot \exp(u) $ and namely for its series expression when $u$ <em>and</em> $v (=u^h) $ are both kept as indeterminate parameter. Pari/GP is still able to evaluate the last expression symbolically and arrives at the same set of coefficients as given in eq (6) in the OP. <hr> <strong>Additional computation:</strong> a short recalculation gives also the leading terms of the powerseries of the following coefficients in (5) of the OP, which I index here as $c_{\lambda,k}$ : </p> <p>$$\small \begin{array} {r} c_{\lambda,1}=&amp; - 1 u^1/1!&amp; - 1 u^2/2!&amp; + 2 u^3/3!&amp; + 11 u^4/4!&amp; + 44 u^5/5!&amp; + 89 u^6/6!&amp; + 636 u^7/7!&amp; + O(u^8) \\ c_{\lambda,2}=&amp;&amp;&amp; - 3 u^3/3!&amp; - 12 u^4/4!&amp; - 5 u^5/5!&amp; + 150 u^6/6!&amp; + 1463 u^7/7!&amp; + O(u^8) \\ c_{\lambda,3}=&amp;&amp;&amp;&amp;&amp; - 40 u^5/5!&amp; - 240 u^6/6!&amp; - 840 u^7/7!&amp; + O(u^8) \\ c_{\lambda,4}=&amp;&amp;&amp;&amp;&amp;&amp;&amp; - 1260 u^7/7!&amp; + O(u^8) \\ \end{array} $$ and recalling that $\; ^hb $ is my notation for your $ \; ^na $ we get $$ \; ^na = \; ^hb = t +c_{\lambda,1} u^h + c_{\lambda,2} u^{2h} + c_{\lambda,3} u^{3h} + \cdots \tag 5$$ Remark: I don't think that a closing term $O(u^{4h})$ makes really sense here because we assume $h&gt;-2$ to be of <em>any real value</em>, especially fractional, and I'm not firm with using that notation in such cases.</p> Answer 5: <p>Let $$c_n(\lambda)=\frac{e^\lambda-{^n a}}{ \lambda^n},\quad d_n(\lambda)=e^{-\lambda}c_n(\lambda)$$ ($d_n$ seems to be simpler). In particular initial functions $c_0(\lambda)=e^\lambda-1$ and $d_0(\lambda)=1-e^{-\lambda}\sim\lambda$ (as $\lambda\to 0$) have good power series of the variable $\lambda$. We are intersted in the limits $$c_\lambda=\lim\limits_{n\to\infty}c_n(\lambda),\quad d_\lambda=\lim\limits_{n\to\infty}d_n(\lambda).$$ The recurrence \begin{gather*} d_{n+1}(\lambda)=\frac{1-e^{-\lambda^{n+1}d_n(\lambda)}}{\lambda^{n+1} } \end{gather*} follows from the definition. By induction $d_n(\lambda)\sim\lambda$. Expanding the recurrence above we get \begin{gather*} d_{n+1}(\lambda)=d_n(\lambda)-\frac{\lambda^{n+1}d_n^2(\lambda)}{2}+\ldots\equiv d_n(\lambda)\pmod{\lambda^{n+3} }. \end{gather*} It means that the limit $d_\lambda$ (as a power series) does exist.</p> <p><strong>EDT.</strong> Numerators $a(k)$ of $c_\lambda$-coefficients are not very random. At least first $41$ coefficients from OP satisfy the congruences \begin{gather*} a(k)\equiv\begin{cases} 0,&amp;\text{if } k\equiv 1\pmod{2 },\\ 1,&amp;\text{if }k\equiv 0\pmod{4 },\\ 3,&amp;\text{if }k\equiv 2\pmod{4 },\\ \end{cases}\pmod{4 }\quad(k\ge 4); \\ a(k)\equiv\begin{cases} 0,&amp;\text{if }k\equiv 1,2\pmod{6 },\\ 1,&amp;\text{if }k\equiv0,3,4,5\pmod{6 },\\ \end{cases}\pmod{3 }\quad(k\ge 3). \end{gather*}</p> Answer 6: <p><strong><em>This is not an answer, only an extended comment at V. Reshetnikov's request in his comment <a href="https://mathoverflow.net/a/272157/7710">at my answer</a></em></strong> </p> <p>@VladimirReshetnikov - not so far now. I think this problem of a closed form is also intimately related to the so far unsolved question of the "mag"-numbers as discussed in <a href="https://mathoverflow.net/questions/57627">"Is there an “elegant” non-recursive formula for these coefficients?"</a> which is focusing the series expansion for the Schröder-function $\sigma()$ as I've used it here.</p> <p>Also your approach using q-binomials for a series <a href="https://mathoverflow.net/questions/259278">"An explicit series representation for the analytic tetration with complex height"</a> was/is much interesting for me - but although I had similar analyses involving q-binomials I didn't arrive at any conciser formulae than the formal algebraic descriptions following from the matrix-based concepts as discussed, for instance, here: <a href="http://go.helms-net.de/math/tetdocs/ExponentialPolynomialInterpolation.pdf" rel="nofollow noreferrer">"Exponential polynomial interpolation"</a> and also here <a href="http://go.helms-net.de/math/tetdocs/BinomialDiagonalization.htm" rel="nofollow noreferrer">"Relation between binomial- and diagonalization method"</a> but at least showing at its end the series <strong><em>(5)</em></strong> in your OP assuming $\lambda=\log 2$ with evaluated coefficients in real numbers and which gives also a short resume of the -as I christened it- "exponential polynomial interpolation" method which involves the q-binomials.</p> <p>In general: the -in my opinion- most sophisticated explication of the coefficients of the function $t^x-1$ (and by conjugacy as well of $b^x$ ) in terms of the logarithm $u$ of $t$ and of $u^h$ comes via the diagonalization-approach as shown in <a href="http://go.helms-net.de/math/tetdocs/APT.htm" rel="nofollow noreferrer">"Eigendecomposition of triangular matrix-operators (here: Ut)"</a> from where the coefficients of your formulae <strong><em>(5)</em></strong> and <strong><em>(6)</em></strong> can be derived, and even symbolically<br> But as well as with the q-binomial approach I did not yet find any simplifications/closed forms besides that (recursive) basic generating schemes for that formulae/coefficients, I am really sorry. </p> <p>Of course, in case I'll find something worth sometime I'll surely link to that discussion here.</p> Answer 7: <p>The coefficients of the series (6) can be found symbolically in $\lambda$ and rational numbers. </p> <p>Consider some base $b$ with $1 \le b \le e^{1/e}$ , the fixpoint $t$ for $f(x)=b^x $ and $f(t)=t$. For bases $b$ in this interval we have a real fixpoint $t$ with $1 \le t \le e$ and $\lambda = \log(t) $ with $0 \le \lambda \le 1$ with $b^t = t$ <em>(which is also attracting when $\lambda,t,b$ are not on the bounds of the resp. intervals)</em>. </p> <p>$\ $ </p> <p>We assume now a conjugate function $d(z)=z/t-1$ and $d°^{-1}(z)=(z+1)\cdot t$ which I write here for shortness in sup-suffix notation $z´ = d(z) $ and $z`=d°^{-1}(z)$ .</p> <p>The <strong><em><a href="https://en.wikipedia.org/wiki/Schr%C3%B6der%27s_equation" rel="nofollow noreferrer">Schröder-mechanism</a></em></strong>, using this conjugacy, allows an invertible "<em>schlicht</em>" function $\sigma(x)$ such that $$ f°^h(x) = \left(\sigma°^{-1}( \lambda^h \cdot \sigma(x´))\right)`$$ </p> <p>Now the given problem assumes $x=1$ such that $x´ = 1/t-1 = \exp(-\lambda)-1$ and we have explicitely $$ f°^h(1) = \left(\sigma°^{-1}( \lambda^h \cdot \sigma(\exp(-\lambda)-1))\right) \cdot \exp(\lambda) + t$$ and because $\sigma(x) = x + O(x^2) $ (meaning to be called "<em>schlicht</em>" ) also its formal power series inverse is of this form and we get $$ f°^h(1) = t + \lambda^h c_\lambda + O(\lambda^{2h})$$ where $c_\lambda = \sigma(1/t-1) $<br> Interestingly, leaving $\lambda$ indeterminate and $t=\exp(\lambda)$ unevaluated as formal power series Pari/GP gives $c_\lambda$ as formal power series in the form as the OP in def. (6) provides its leading coefficients. </p> <hr> <blockquote> <p><strong>Derivation/implementation in Pari/GP</strong> </p> </blockquote> <p>Here I show, illustrated by sample calculations with Pari/GP, that there is no numerical approximation needed to get the coefficients at $\lambda^k$ in definition (6). Instead there is a derivation on formal power series which provides exact rational coefficients. The method below reproduces perfectly your finding. </p> <p><hr> <strong>Preliminaries:</strong> a <a href="https://en.wikipedia.org/wiki/Carleman_matrix" rel="nofollow noreferrer">Carleman-matrix</a> <strong><em>B</em></strong> associated to a function $f(x)$ which has a power series with more than zero radius of convergence (and generalized: can be Euler-summed in a certain interval of its argument) is initially only a collection of column-vectors $B[,c]$ <em>($c$ indexing the c-olumn)</em>, which contain the coefficients of the formal power series of $f(x)$ and more precisely of $f(x)^c$ for $c=0 \ldots \infty$ in the column of index $c$ , and is thus in both direction of infinite size. It is an important special case when the power series of some function $f(x)$ has no constant term: then the shape of $B$ is lower triangular and the analytic handling of such Carlemanmatrices has been substantially established. We'll need this in the derivation below.<br> <sub> It is not trivial to show, that such a <em>collection</em> of columns-vectors can indeed be used as <strong><em>Matrix</em></strong>, applicable to concepts of matrix-algebra and -analysis, for instance raising to powers and especially diagonalization - but there is reliable literature about this so that we can omit such a step here. One pioneer is surely Eri Jabotinsky, see for instance this 1963-article <a href="http://www.jstor.org/stable/1993593" rel="nofollow noreferrer">"Analytic iteration"</a> (introducing the term "representation matrix")</sub> </p> <p>After the Carlemanmatrix <strong><em>B</em></strong> contains the coefficients of formal power series we need also a construct $V(x)$ as row vector $V(x)=[1,x,x^2,x^3,\ldots]$ which I call of "Vandermonde"-type here. Using $V(x)$ as diagonal-matrix I pre-superscribe this like $ \;^dV(x)$ <hr> For instance the Carlemanmatrix $B_e$ for the function $f(x) = \exp(x)$ looks like</p> <pre><code> 1 1 1 1 1 ... 0 1 2 3 4 ... 0 1/2! 2^2/2! 3^2/2! 4^2/2! ... 0 1/3! 2^3/3! 3^3/3! 4^3/3! ... 0 1/4! 2^4/4! 3^4/4! 4^4/4! ... ... ... ... ... .... ... </code></pre> <p>The clue with such Carlemanmatrices is, that the left-multiplication of some $B$ with a vector $V(x)$ as in $$V(x) \cdot B = V(f(x)) \tag 1$$ gives a resulting vector of the same structure $V(f(x)) = [ 1 , f(x), f(x)^2, f(x)^3, ... ]$ .<br> Of course, using the only "relevant" column in $B$ alone , then this gives the scalar expression $f(x)$: $$V(x) \cdot B [,1] = f(x) \tag {1.1}$$ </p> <p>This means: a composition of a function with itself $f(f(x))$ (in terms of its power series) is mapped to a <em>repeated multiplication</em> with its Carleman-matrix or in other words: is mapped to multiplication with <strong><em>powers</em></strong> of its Carlemanmatrix. </p> <p>The further clue is now, that for appropriate matrices a <strong><em>fractional</em></strong> power can be determined and for that cases we can thus easily define power series for the fractional iterates of the associated function. </p> <hr> <p><strong>Technically</strong> this means that the concept of diagonalization can be transferred to that Carleman-ansatz (of course for suitable Carlemanmatrices $B$). A case, for which this is widely accepted is the case where the Carleman-matrices for some function $f(x)$ is lower triangular.</p> <p>We have the function $f(x) = b^x $ and by the definition in the OP $1 \lt b \lt e^{1/e}$ . With this restriction, $f(x)$ has a fixpoint $t$ such that $f(t)=t$ (or: $b^t=t$ ) and also this fixpoint is attracting (which is important here). We can, as I wrote in the first part of text above, apply conjugacy to arrive at the function $g(x)=f(x`)´ = t^x-1 = ux + (ux)^2/2! + O(x^3) $ (for notational convenience I write $u$ instead of $\lambda$ in the following). </p> <p>Given the lower triangular matrix of Stirlingnumbers $2$'nd kind $S2$ and its factorially similarity-scaled transform which I denote here as $fS2F$ we have the lower triangular Carlemanmatrix $U_t$ for $g(x)$ by $$ U_t = \; ^dV(u) \cdot fS2F \tag {2.1}$$</p> <hr> <p>The top-left edge of $U_t$ looks like</p> <pre><code> 1 . . . . . 0 u/1! . . . . 0 u^2/2! u^2*2!/2! . . . 0 u^3/3! 3*u^3*2!/3! u^3*3!/3! . . 0 u^4/4! 7*u^4*2!/4! 6*u^4*3!/4! u^4*4!/4! . 0 u^5/5! 15*u^5*2!/5! 25*u^5*3!/5! 10*u^5*4!/5! u^5*5!/5! </code></pre> <p>or with cancelled entries </p> <pre><code> 1 . . . . . 0 u . . . . 0 1/2*u^2 u^2 . . . 0 1/6*u^3 u^3 u^3 . . 0 1/24*u^4 7/12*u^4 3/2*u^4 u^4 . 0 1/120*u^5 1/4*u^5 5/4*u^5 2*u^5 u^5 </code></pre> <p>and taking the coefficients from the second column $$ g(x) = \sum_{k=1}^\infty u^k \cdot \frac{x^k}{ k!} \tag {2.2}$$ <hr> The formal power series for $h$'th fractional iterates $g°^h(x)$ are now generated by fractional powers of $U_t$ instead and this can be done by diagonalization: </p> <p>$$ U_t = M \cdot D \cdot W \qquad \qquad \text{where I denote $W$ for } M^{-1} \tag {3.1}$$ where also the diagonal entries in $D$ are the consecutive powers of $u$ $$ D = \; ^d V(u) \text{ and } D^h = \; ^dV(u^h) \tag {3.2}$$ such that $$ U_t^h = M \cdot \; ^dV(u^h) \cdot W \tag {3.3} $$</p> <p>Unfortunately the current Pari/GP-version's diagonalization procedure cannot exploit the advantage of $U_t$ being triangular and also cannot keep the parameter $u$ symbolically. But it is not difficult to have a small recursive routine to generate $M$ , $D$ and $W (=M^{-1})$ for the triangular Carleman-case; the additional norming of $M$ to have a unit-diagonal is also included and with this the matrix $M$ is also of the Carleman-type (for details and an implementation in Pari/GP see the explanation in my linked article). The matrix <strong><em>M</em></strong> is now the Carlemanmatrix for the above introduced Schröder-function $\sigma(x)$ which takes its coefficients from the second column in <strong><em>M</em></strong>. </p> <p>The top-left of the matrix $M$ looks like</p> <pre><code> 1 . . . 0 1 . . 0 u/(-2*u+2) 1 . 0 (2*u^3+u^2)/(6*u^3-6*u^2-6*u+6) u/(-u+1) 1 </code></pre> <p>The entries in the (only relevant) $2$'nd column have a partially discernable pattern; I've described this in earlier discussions, see for instance <a href="http://go.helms-net.de/math/tetdocs/APT.htm" rel="nofollow noreferrer">here</a>, but which is too much to show at this place.<br> <hr> Having this we can write for the $h$'th fractional iterate $g(x)$ of some $x$: </p> <p>$$ \begin{array}{rll} V(x) \cdot U_t^h &amp;= V( g°^h(x)) \\ &amp;= V(x) \cdot (M \cdot \;^dV(u^h) \cdot W) \qquad \qquad \text{ and by associativity}\\ &amp;= (V(x) \cdot M) \cdot \;^dV(u^h) \cdot W \end{array} \tag {4.1}$$</p> <p>Because $M$ is of Carlemantype we have also $$ V(x) \cdot M = V(\sigma(x)) \tag {4.2}$$ or -using also conjugacy- $$ V(x´) \cdot M = V(\sigma(x´)) \tag {4.3}$$ and it is relevant here, that $\sigma(x´)$ is <strong><em>identical to the (normed) Schröder-function</em></strong> for the iteration of $b^x$. </p> <p>In the problem as given in the OP we assume $x=1$ or $x´=1/t-1 = \exp(-u)-1$ and to have the Schröder-function symbolically we have to look at $$ V(1´) \cdot M = V(\exp(-u)-1) \cdot M = V(\sigma(\exp(-u)-1)) \tag {4.4} $$ and this can still be done symbolically in $u$; having <strong><em>M</em></strong> in its symbolical representation, Pari/GP is able to give the leading part of $\sigma(\exp(-u)-1)$ by formal expansion of $\exp(-u)$ as power series taken as argument of $\sigma(\cdot)$ with still rational coefficients. </p> <p>Having the left part of $(4.2)$ evaluated we need now $$ V(\sigma(\exp(-u)-1)) \cdot \; ^dV(u^h) \cdot W[,1] = y =\sigma°^{-1}(u^h \cdot \sigma(1´))\tag {4.5}$$ by the relevant $2$'nd column of $W$ only. </p> <p>Note that this gives a power series in $u^h$ with the (constant and rational) coefficients in $W[,1]$ and powers of the expression $\sigma(\exp(-u)-1)$ . To help Pari/GP to keep $u^h$ as a single symbol, we need to substitute it by the name of a unevaluated variable, say $u_h$ or $v$ and re-substitute it later after series-expansions.</p> <p>The final result $f°^h(1) $ occurs by the inverse conjugacy on $y$ $$f°^h(1) = y` = (y+1)\cdot t = y\cdot t + t =t + y \cdot \exp(u) \tag {4.6} $$ which can still be expressed as formal power series in the indeterminate $u$ and $u^h$. </p> <p>The OP asks now for the first coefficient in $y \cdot \exp(u) $ and namely for its series expression when $u$ <em>and</em> $v (=u^h) $ are both kept as indeterminate parameter. Pari/GP is still able to evaluate the last expression symbolically and arrives at the same set of coefficients as given in eq (6) in the OP. <hr> <strong>Additional computation:</strong> a short recalculation gives also the leading terms of the powerseries of the following coefficients in (5) of the OP, which I index here as $c_{\lambda,k}$ : </p> <p>$$\small \begin{array} {r} c_{\lambda,1}=&amp; - 1 u^1/1!&amp; - 1 u^2/2!&amp; + 2 u^3/3!&amp; + 11 u^4/4!&amp; + 44 u^5/5!&amp; + 89 u^6/6!&amp; + 636 u^7/7!&amp; + O(u^8) \\ c_{\lambda,2}=&amp;&amp;&amp; - 3 u^3/3!&amp; - 12 u^4/4!&amp; - 5 u^5/5!&amp; + 150 u^6/6!&amp; + 1463 u^7/7!&amp; + O(u^8) \\ c_{\lambda,3}=&amp;&amp;&amp;&amp;&amp; - 40 u^5/5!&amp; - 240 u^6/6!&amp; - 840 u^7/7!&amp; + O(u^8) \\ c_{\lambda,4}=&amp;&amp;&amp;&amp;&amp;&amp;&amp; - 1260 u^7/7!&amp; + O(u^8) \\ \end{array} $$ and recalling that $\; ^hb $ is my notation for your $ \; ^na $ we get $$ \; ^na = \; ^hb = t +c_{\lambda,1} u^h + c_{\lambda,2} u^{2h} + c_{\lambda,3} u^{3h} + \cdots \tag 5$$ Remark: I don't think that a closing term $O(u^{4h})$ makes really sense here because we assume $h&gt;-2$ to be of <em>any real value</em>, especially fractional, and I'm not firm with using that notation in such cases.</p> Answer 8: <p>Let $$c_n(\lambda)=\frac{e^\lambda-{^n a}}{ \lambda^n},\quad d_n(\lambda)=e^{-\lambda}c_n(\lambda)$$ ($d_n$ seems to be simpler). In particular initial functions $c_0(\lambda)=e^\lambda-1$ and $d_0(\lambda)=1-e^{-\lambda}\sim\lambda$ (as $\lambda\to 0$) have good power series of the variable $\lambda$. We are intersted in the limits $$c_\lambda=\lim\limits_{n\to\infty}c_n(\lambda),\quad d_\lambda=\lim\limits_{n\to\infty}d_n(\lambda).$$ The recurrence \begin{gather*} d_{n+1}(\lambda)=\frac{1-e^{-\lambda^{n+1}d_n(\lambda)}}{\lambda^{n+1} } \end{gather*} follows from the definition. By induction $d_n(\lambda)\sim\lambda$. Expanding the recurrence above we get \begin{gather*} d_{n+1}(\lambda)=d_n(\lambda)-\frac{\lambda^{n+1}d_n^2(\lambda)}{2}+\ldots\equiv d_n(\lambda)\pmod{\lambda^{n+3} }. \end{gather*} It means that the limit $d_\lambda$ (as a power series) does exist.</p> <p><strong>EDT.</strong> Numerators $a(k)$ of $c_\lambda$-coefficients are not very random. At least first $41$ coefficients from OP satisfy the congruences \begin{gather*} a(k)\equiv\begin{cases} 0,&amp;\text{if } k\equiv 1\pmod{2 },\\ 1,&amp;\text{if }k\equiv 0\pmod{4 },\\ 3,&amp;\text{if }k\equiv 2\pmod{4 },\\ \end{cases}\pmod{4 }\quad(k\ge 4); \\ a(k)\equiv\begin{cases} 0,&amp;\text{if }k\equiv 1,2\pmod{6 },\\ 1,&amp;\text{if }k\equiv0,3,4,5\pmod{6 },\\ \end{cases}\pmod{3 }\quad(k\ge 3). \end{gather*}</p> Answer 9: <p><strong><em>This is not an answer, only an extended comment at V. Reshetnikov's request in his comment <a href="https://mathoverflow.net/a/272157/7710">at my answer</a></em></strong> </p> <p>@VladimirReshetnikov - not so far now. I think this problem of a closed form is also intimately related to the so far unsolved question of the "mag"-numbers as discussed in <a href="https://mathoverflow.net/questions/57627">"Is there an “elegant” non-recursive formula for these coefficients?"</a> which is focusing the series expansion for the Schröder-function $\sigma()$ as I've used it here.</p> <p>Also your approach using q-binomials for a series <a href="https://mathoverflow.net/questions/259278">"An explicit series representation for the analytic tetration with complex height"</a> was/is much interesting for me - but although I had similar analyses involving q-binomials I didn't arrive at any conciser formulae than the formal algebraic descriptions following from the matrix-based concepts as discussed, for instance, here: <a href="http://go.helms-net.de/math/tetdocs/ExponentialPolynomialInterpolation.pdf" rel="nofollow noreferrer">"Exponential polynomial interpolation"</a> and also here <a href="http://go.helms-net.de/math/tetdocs/BinomialDiagonalization.htm" rel="nofollow noreferrer">"Relation between binomial- and diagonalization method"</a> but at least showing at its end the series <strong><em>(5)</em></strong> in your OP assuming $\lambda=\log 2$ with evaluated coefficients in real numbers and which gives also a short resume of the -as I christened it- "exponential polynomial interpolation" method which involves the q-binomials.</p> <p>In general: the -in my opinion- most sophisticated explication of the coefficients of the function $t^x-1$ (and by conjugacy as well of $b^x$ ) in terms of the logarithm $u$ of $t$ and of $u^h$ comes via the diagonalization-approach as shown in <a href="http://go.helms-net.de/math/tetdocs/APT.htm" rel="nofollow noreferrer">"Eigendecomposition of triangular matrix-operators (here: Ut)"</a> from where the coefficients of your formulae <strong><em>(5)</em></strong> and <strong><em>(6)</em></strong> can be derived, and even symbolically<br> But as well as with the q-binomial approach I did not yet find any simplifications/closed forms besides that (recursive) basic generating schemes for that formulae/coefficients, I am really sorry. </p> <p>Of course, in case I'll find something worth sometime I'll surely link to that discussion here.</p>
Is the Hodge Conjecture an $\mathbb{A}^1$-homotopy invariant?
260,671
18
https://mathoverflow.net/questions/260671/is-the-hodge-conjecture-an-mathbba1-homotopy-invariant
<p>Let $X$ and $Y$ be two nonsingular projective varieties defined over the complex numbers. </p> <blockquote> <p>If $X$ and $Y$ are $\mathbb{A}^1$-homotopy equivalent, then does $X$ satisfying the Hodge conjecture imply that $Y$ satisfies the Hodge conjecture?</p> </blockquote>
[ { "answer_id": 260920, "score": 15, "text": "<p>Isn't this very easy? If the varieties are $\\mathbb{A}^1$-homotopy equivalent, then their Voevodsky motives are isomorphic also (since there is a connecting functor making the obvious diagram commutative). So it remains to note that all the ingredients of your question are \"motivic\". For this purpose one may recall that Chow motives embed into Voevodsky one and note that both the cycle classes and the Hodge classes are determined by Chow motives of the corresponding varieties. </p>\n" }, { "answer_id": 260758, "score": 5, "text": "<p>This is more of a very long comment, but I believe so.</p>\n\n<p>A modern description of the motivic homotopy category is $sPre(Sm/k)[N^{-1}][L^{-1}]$ where $N^{-1}$ is localizing with respect to the topology and $L^{-1}$ is localizing with respect to all the maps $X \\times \\mathbb{A}^1 \\rightarrow X$. Chow groups of course satisfy Nisnevich descent, and are $\\mathbb{A}^1$ invariant, and so they give a motivic sheaf. The question is then if the Hodge classes can be extended to a motivic sheaf. Since the definition of Hodge classes is only for smooth and projective $X$, it's not totally obvious what to use for a general smooth variety.</p>\n\n<p>Here we should use the fact that Betti realization $X \\mapsto X^{an}$ is a motivic sheaf of spaces, and that both the weight-filtered , I'll write $\\hat{\\mathcal{W}_k}A_{\\log}$, and also $\\mathcal{F}^p \\cap \\hat{\\mathcal{W}_k}A_{\\log}$ the weight-and-Hodge-filtered de Rham complexes are perfectly good motivic sheaves of cochain complexes (up to q.iso). Here I'm using the notation of Bunke and Tamme from \"Regulators and cycle maps...\", where they define them carefully. Bunke and Tamme prove Zariski descent for the latter two in Lemma 3.6, but their method actually shows etale descent. $\\mathbb{A}^1$-invariance can be checked on the level of cohomology. We can just take the set-theoretic conjugate to get a sheaf $\\overline{\\mathcal{F}^p} \\cap \\hat{\\mathcal{W}_k}A_{\\log}$. To just work with sheaves of cochain complexes (up to q.iso), let's look at the singular cochains functor instead of the Betti realization functor. That is, let's look at $X \\mapsto Sing_\\mathbb{Q}^\\bullet(X)$. We have a diagram</p>\n\n<p>$\\require{AMScd}$\n\\begin{CD}\n \\mathcal{F}^p \\cap \\hat{\\mathcal{W}_{2p}}A_{\\log} @&gt;&gt;&gt; A_{\\log} @&lt;&lt;&lt; \\overline{\\mathcal{F}^p} \\cap \\hat{\\mathcal{W}_{2p}}A_{\\log} \\\\\n @. @AAA \\\\\n@. Sing_\\mathbb{Q}^\\bullet(X)\n\\end{CD}</p>\n\n<p>Whose homotopy limit sheaf should output the $p$-Hodge classes for input a smooth projective $X$.</p>\n" }, { "answer_id": 260760, "score": 4, "text": "<p>I was initially reluctant to contribute an answer because I'm not an expert on $\\mathbb{A}^1$ homotopy theory. But let me flesh out my first comment a little, and you can decide whether this is close to what you wanted. Let $K_0(Var)$ be the Grothendieck ring of all complex algebraic varieties. Let $\\mathbb{L}$ denote the class of the affine line. The quotient $K_0(Var)/(\\mathbb{L}-1)$ might be thought of as the Grothendieck ring of <em>naive</em> $\\mathbb{A}^1$-homotopy classes. This maps to $K_0(Var)[\\mathbb{L}^{-1}]$ and the Hodge conjecture \"factors through\" the second group by <a href=\"http://arxiv.org/abs/math/0506210\" rel=\"noreferrer\">http://arxiv.org/abs/math/0506210</a></p>\n" }, { "answer_id": 260920, "score": 15, "text": "<p>Isn't this very easy? If the varieties are $\\mathbb{A}^1$-homotopy equivalent, then their Voevodsky motives are isomorphic also (since there is a connecting functor making the obvious diagram commutative). So it remains to note that all the ingredients of your question are \"motivic\". For this purpose one may recall that Chow motives embed into Voevodsky one and note that both the cycle classes and the Hodge classes are determined by Chow motives of the corresponding varieties. </p>\n" }, { "answer_id": 260758, "score": 5, "text": "<p>This is more of a very long comment, but I believe so.</p>\n\n<p>A modern description of the motivic homotopy category is $sPre(Sm/k)[N^{-1}][L^{-1}]$ where $N^{-1}$ is localizing with respect to the topology and $L^{-1}$ is localizing with respect to all the maps $X \\times \\mathbb{A}^1 \\rightarrow X$. Chow groups of course satisfy Nisnevich descent, and are $\\mathbb{A}^1$ invariant, and so they give a motivic sheaf. The question is then if the Hodge classes can be extended to a motivic sheaf. Since the definition of Hodge classes is only for smooth and projective $X$, it's not totally obvious what to use for a general smooth variety.</p>\n\n<p>Here we should use the fact that Betti realization $X \\mapsto X^{an}$ is a motivic sheaf of spaces, and that both the weight-filtered , I'll write $\\hat{\\mathcal{W}_k}A_{\\log}$, and also $\\mathcal{F}^p \\cap \\hat{\\mathcal{W}_k}A_{\\log}$ the weight-and-Hodge-filtered de Rham complexes are perfectly good motivic sheaves of cochain complexes (up to q.iso). Here I'm using the notation of Bunke and Tamme from \"Regulators and cycle maps...\", where they define them carefully. Bunke and Tamme prove Zariski descent for the latter two in Lemma 3.6, but their method actually shows etale descent. $\\mathbb{A}^1$-invariance can be checked on the level of cohomology. We can just take the set-theoretic conjugate to get a sheaf $\\overline{\\mathcal{F}^p} \\cap \\hat{\\mathcal{W}_k}A_{\\log}$. To just work with sheaves of cochain complexes (up to q.iso), let's look at the singular cochains functor instead of the Betti realization functor. That is, let's look at $X \\mapsto Sing_\\mathbb{Q}^\\bullet(X)$. We have a diagram</p>\n\n<p>$\\require{AMScd}$\n\\begin{CD}\n \\mathcal{F}^p \\cap \\hat{\\mathcal{W}_{2p}}A_{\\log} @&gt;&gt;&gt; A_{\\log} @&lt;&lt;&lt; \\overline{\\mathcal{F}^p} \\cap \\hat{\\mathcal{W}_{2p}}A_{\\log} \\\\\n @. @AAA \\\\\n@. Sing_\\mathbb{Q}^\\bullet(X)\n\\end{CD}</p>\n\n<p>Whose homotopy limit sheaf should output the $p$-Hodge classes for input a smooth projective $X$.</p>\n" }, { "answer_id": 260760, "score": 4, "text": "<p>I was initially reluctant to contribute an answer because I'm not an expert on $\\mathbb{A}^1$ homotopy theory. But let me flesh out my first comment a little, and you can decide whether this is close to what you wanted. Let $K_0(Var)$ be the Grothendieck ring of all complex algebraic varieties. Let $\\mathbb{L}$ denote the class of the affine line. The quotient $K_0(Var)/(\\mathbb{L}-1)$ might be thought of as the Grothendieck ring of <em>naive</em> $\\mathbb{A}^1$-homotopy classes. This maps to $K_0(Var)[\\mathbb{L}^{-1}]$ and the Hodge conjecture \"factors through\" the second group by <a href=\"http://arxiv.org/abs/math/0506210\" rel=\"noreferrer\">http://arxiv.org/abs/math/0506210</a></p>\n" }, { "answer_id": 260920, "score": 15, "text": "<p>Isn't this very easy? If the varieties are $\\mathbb{A}^1$-homotopy equivalent, then their Voevodsky motives are isomorphic also (since there is a connecting functor making the obvious diagram commutative). So it remains to note that all the ingredients of your question are \"motivic\". For this purpose one may recall that Chow motives embed into Voevodsky one and note that both the cycle classes and the Hodge classes are determined by Chow motives of the corresponding varieties. </p>\n" }, { "answer_id": 260758, "score": 5, "text": "<p>This is more of a very long comment, but I believe so.</p>\n\n<p>A modern description of the motivic homotopy category is $sPre(Sm/k)[N^{-1}][L^{-1}]$ where $N^{-1}$ is localizing with respect to the topology and $L^{-1}$ is localizing with respect to all the maps $X \\times \\mathbb{A}^1 \\rightarrow X$. Chow groups of course satisfy Nisnevich descent, and are $\\mathbb{A}^1$ invariant, and so they give a motivic sheaf. The question is then if the Hodge classes can be extended to a motivic sheaf. Since the definition of Hodge classes is only for smooth and projective $X$, it's not totally obvious what to use for a general smooth variety.</p>\n\n<p>Here we should use the fact that Betti realization $X \\mapsto X^{an}$ is a motivic sheaf of spaces, and that both the weight-filtered , I'll write $\\hat{\\mathcal{W}_k}A_{\\log}$, and also $\\mathcal{F}^p \\cap \\hat{\\mathcal{W}_k}A_{\\log}$ the weight-and-Hodge-filtered de Rham complexes are perfectly good motivic sheaves of cochain complexes (up to q.iso). Here I'm using the notation of Bunke and Tamme from \"Regulators and cycle maps...\", where they define them carefully. Bunke and Tamme prove Zariski descent for the latter two in Lemma 3.6, but their method actually shows etale descent. $\\mathbb{A}^1$-invariance can be checked on the level of cohomology. We can just take the set-theoretic conjugate to get a sheaf $\\overline{\\mathcal{F}^p} \\cap \\hat{\\mathcal{W}_k}A_{\\log}$. To just work with sheaves of cochain complexes (up to q.iso), let's look at the singular cochains functor instead of the Betti realization functor. That is, let's look at $X \\mapsto Sing_\\mathbb{Q}^\\bullet(X)$. We have a diagram</p>\n\n<p>$\\require{AMScd}$\n\\begin{CD}\n \\mathcal{F}^p \\cap \\hat{\\mathcal{W}_{2p}}A_{\\log} @&gt;&gt;&gt; A_{\\log} @&lt;&lt;&lt; \\overline{\\mathcal{F}^p} \\cap \\hat{\\mathcal{W}_{2p}}A_{\\log} \\\\\n @. @AAA \\\\\n@. Sing_\\mathbb{Q}^\\bullet(X)\n\\end{CD}</p>\n\n<p>Whose homotopy limit sheaf should output the $p$-Hodge classes for input a smooth projective $X$.</p>\n" }, { "answer_id": 260760, "score": 4, "text": "<p>I was initially reluctant to contribute an answer because I'm not an expert on $\\mathbb{A}^1$ homotopy theory. But let me flesh out my first comment a little, and you can decide whether this is close to what you wanted. Let $K_0(Var)$ be the Grothendieck ring of all complex algebraic varieties. Let $\\mathbb{L}$ denote the class of the affine line. The quotient $K_0(Var)/(\\mathbb{L}-1)$ might be thought of as the Grothendieck ring of <em>naive</em> $\\mathbb{A}^1$-homotopy classes. This maps to $K_0(Var)[\\mathbb{L}^{-1}]$ and the Hodge conjecture \"factors through\" the second group by <a href=\"http://arxiv.org/abs/math/0506210\" rel=\"noreferrer\">http://arxiv.org/abs/math/0506210</a></p>\n" } ]
Question: Is the Hodge Conjecture an $\mathbb{A}^1$-homotopy invariant? Answer 1: <p>Isn't this very easy? If the varieties are $\mathbb{A}^1$-homotopy equivalent, then their Voevodsky motives are isomorphic also (since there is a connecting functor making the obvious diagram commutative). So it remains to note that all the ingredients of your question are "motivic". For this purpose one may recall that Chow motives embed into Voevodsky one and note that both the cycle classes and the Hodge classes are determined by Chow motives of the corresponding varieties. </p> Answer 2: <p>This is more of a very long comment, but I believe so.</p> <p>A modern description of the motivic homotopy category is $sPre(Sm/k)[N^{-1}][L^{-1}]$ where $N^{-1}$ is localizing with respect to the topology and $L^{-1}$ is localizing with respect to all the maps $X \times \mathbb{A}^1 \rightarrow X$. Chow groups of course satisfy Nisnevich descent, and are $\mathbb{A}^1$ invariant, and so they give a motivic sheaf. The question is then if the Hodge classes can be extended to a motivic sheaf. Since the definition of Hodge classes is only for smooth and projective $X$, it's not totally obvious what to use for a general smooth variety.</p> <p>Here we should use the fact that Betti realization $X \mapsto X^{an}$ is a motivic sheaf of spaces, and that both the weight-filtered , I'll write $\hat{\mathcal{W}_k}A_{\log}$, and also $\mathcal{F}^p \cap \hat{\mathcal{W}_k}A_{\log}$ the weight-and-Hodge-filtered de Rham complexes are perfectly good motivic sheaves of cochain complexes (up to q.iso). Here I'm using the notation of Bunke and Tamme from "Regulators and cycle maps...", where they define them carefully. Bunke and Tamme prove Zariski descent for the latter two in Lemma 3.6, but their method actually shows etale descent. $\mathbb{A}^1$-invariance can be checked on the level of cohomology. We can just take the set-theoretic conjugate to get a sheaf $\overline{\mathcal{F}^p} \cap \hat{\mathcal{W}_k}A_{\log}$. To just work with sheaves of cochain complexes (up to q.iso), let's look at the singular cochains functor instead of the Betti realization functor. That is, let's look at $X \mapsto Sing_\mathbb{Q}^\bullet(X)$. We have a diagram</p> <p>$\require{AMScd}$ \begin{CD} \mathcal{F}^p \cap \hat{\mathcal{W}_{2p}}A_{\log} @&gt;&gt;&gt; A_{\log} @&lt;&lt;&lt; \overline{\mathcal{F}^p} \cap \hat{\mathcal{W}_{2p}}A_{\log} \\ @. @AAA \\ @. Sing_\mathbb{Q}^\bullet(X) \end{CD}</p> <p>Whose homotopy limit sheaf should output the $p$-Hodge classes for input a smooth projective $X$.</p> Answer 3: <p>I was initially reluctant to contribute an answer because I'm not an expert on $\mathbb{A}^1$ homotopy theory. But let me flesh out my first comment a little, and you can decide whether this is close to what you wanted. Let $K_0(Var)$ be the Grothendieck ring of all complex algebraic varieties. Let $\mathbb{L}$ denote the class of the affine line. The quotient $K_0(Var)/(\mathbb{L}-1)$ might be thought of as the Grothendieck ring of <em>naive</em> $\mathbb{A}^1$-homotopy classes. This maps to $K_0(Var)[\mathbb{L}^{-1}]$ and the Hodge conjecture "factors through" the second group by <a href="http://arxiv.org/abs/math/0506210" rel="noreferrer">http://arxiv.org/abs/math/0506210</a></p> Answer 4: <p>Isn't this very easy? If the varieties are $\mathbb{A}^1$-homotopy equivalent, then their Voevodsky motives are isomorphic also (since there is a connecting functor making the obvious diagram commutative). So it remains to note that all the ingredients of your question are "motivic". For this purpose one may recall that Chow motives embed into Voevodsky one and note that both the cycle classes and the Hodge classes are determined by Chow motives of the corresponding varieties. </p> Answer 5: <p>This is more of a very long comment, but I believe so.</p> <p>A modern description of the motivic homotopy category is $sPre(Sm/k)[N^{-1}][L^{-1}]$ where $N^{-1}$ is localizing with respect to the topology and $L^{-1}$ is localizing with respect to all the maps $X \times \mathbb{A}^1 \rightarrow X$. Chow groups of course satisfy Nisnevich descent, and are $\mathbb{A}^1$ invariant, and so they give a motivic sheaf. The question is then if the Hodge classes can be extended to a motivic sheaf. Since the definition of Hodge classes is only for smooth and projective $X$, it's not totally obvious what to use for a general smooth variety.</p> <p>Here we should use the fact that Betti realization $X \mapsto X^{an}$ is a motivic sheaf of spaces, and that both the weight-filtered , I'll write $\hat{\mathcal{W}_k}A_{\log}$, and also $\mathcal{F}^p \cap \hat{\mathcal{W}_k}A_{\log}$ the weight-and-Hodge-filtered de Rham complexes are perfectly good motivic sheaves of cochain complexes (up to q.iso). Here I'm using the notation of Bunke and Tamme from "Regulators and cycle maps...", where they define them carefully. Bunke and Tamme prove Zariski descent for the latter two in Lemma 3.6, but their method actually shows etale descent. $\mathbb{A}^1$-invariance can be checked on the level of cohomology. We can just take the set-theoretic conjugate to get a sheaf $\overline{\mathcal{F}^p} \cap \hat{\mathcal{W}_k}A_{\log}$. To just work with sheaves of cochain complexes (up to q.iso), let's look at the singular cochains functor instead of the Betti realization functor. That is, let's look at $X \mapsto Sing_\mathbb{Q}^\bullet(X)$. We have a diagram</p> <p>$\require{AMScd}$ \begin{CD} \mathcal{F}^p \cap \hat{\mathcal{W}_{2p}}A_{\log} @&gt;&gt;&gt; A_{\log} @&lt;&lt;&lt; \overline{\mathcal{F}^p} \cap \hat{\mathcal{W}_{2p}}A_{\log} \\ @. @AAA \\ @. Sing_\mathbb{Q}^\bullet(X) \end{CD}</p> <p>Whose homotopy limit sheaf should output the $p$-Hodge classes for input a smooth projective $X$.</p> Answer 6: <p>I was initially reluctant to contribute an answer because I'm not an expert on $\mathbb{A}^1$ homotopy theory. But let me flesh out my first comment a little, and you can decide whether this is close to what you wanted. Let $K_0(Var)$ be the Grothendieck ring of all complex algebraic varieties. Let $\mathbb{L}$ denote the class of the affine line. The quotient $K_0(Var)/(\mathbb{L}-1)$ might be thought of as the Grothendieck ring of <em>naive</em> $\mathbb{A}^1$-homotopy classes. This maps to $K_0(Var)[\mathbb{L}^{-1}]$ and the Hodge conjecture "factors through" the second group by <a href="http://arxiv.org/abs/math/0506210" rel="noreferrer">http://arxiv.org/abs/math/0506210</a></p> Answer 7: <p>Isn't this very easy? If the varieties are $\mathbb{A}^1$-homotopy equivalent, then their Voevodsky motives are isomorphic also (since there is a connecting functor making the obvious diagram commutative). So it remains to note that all the ingredients of your question are "motivic". For this purpose one may recall that Chow motives embed into Voevodsky one and note that both the cycle classes and the Hodge classes are determined by Chow motives of the corresponding varieties. </p> Answer 8: <p>This is more of a very long comment, but I believe so.</p> <p>A modern description of the motivic homotopy category is $sPre(Sm/k)[N^{-1}][L^{-1}]$ where $N^{-1}$ is localizing with respect to the topology and $L^{-1}$ is localizing with respect to all the maps $X \times \mathbb{A}^1 \rightarrow X$. Chow groups of course satisfy Nisnevich descent, and are $\mathbb{A}^1$ invariant, and so they give a motivic sheaf. The question is then if the Hodge classes can be extended to a motivic sheaf. Since the definition of Hodge classes is only for smooth and projective $X$, it's not totally obvious what to use for a general smooth variety.</p> <p>Here we should use the fact that Betti realization $X \mapsto X^{an}$ is a motivic sheaf of spaces, and that both the weight-filtered , I'll write $\hat{\mathcal{W}_k}A_{\log}$, and also $\mathcal{F}^p \cap \hat{\mathcal{W}_k}A_{\log}$ the weight-and-Hodge-filtered de Rham complexes are perfectly good motivic sheaves of cochain complexes (up to q.iso). Here I'm using the notation of Bunke and Tamme from "Regulators and cycle maps...", where they define them carefully. Bunke and Tamme prove Zariski descent for the latter two in Lemma 3.6, but their method actually shows etale descent. $\mathbb{A}^1$-invariance can be checked on the level of cohomology. We can just take the set-theoretic conjugate to get a sheaf $\overline{\mathcal{F}^p} \cap \hat{\mathcal{W}_k}A_{\log}$. To just work with sheaves of cochain complexes (up to q.iso), let's look at the singular cochains functor instead of the Betti realization functor. That is, let's look at $X \mapsto Sing_\mathbb{Q}^\bullet(X)$. We have a diagram</p> <p>$\require{AMScd}$ \begin{CD} \mathcal{F}^p \cap \hat{\mathcal{W}_{2p}}A_{\log} @&gt;&gt;&gt; A_{\log} @&lt;&lt;&lt; \overline{\mathcal{F}^p} \cap \hat{\mathcal{W}_{2p}}A_{\log} \\ @. @AAA \\ @. Sing_\mathbb{Q}^\bullet(X) \end{CD}</p> <p>Whose homotopy limit sheaf should output the $p$-Hodge classes for input a smooth projective $X$.</p> Answer 9: <p>I was initially reluctant to contribute an answer because I'm not an expert on $\mathbb{A}^1$ homotopy theory. But let me flesh out my first comment a little, and you can decide whether this is close to what you wanted. Let $K_0(Var)$ be the Grothendieck ring of all complex algebraic varieties. Let $\mathbb{L}$ denote the class of the affine line. The quotient $K_0(Var)/(\mathbb{L}-1)$ might be thought of as the Grothendieck ring of <em>naive</em> $\mathbb{A}^1$-homotopy classes. This maps to $K_0(Var)[\mathbb{L}^{-1}]$ and the Hodge conjecture "factors through" the second group by <a href="http://arxiv.org/abs/math/0506210" rel="noreferrer">http://arxiv.org/abs/math/0506210</a></p>
$p$-adic numbers in physics
259,155
17
https://mathoverflow.net/questions/259155/p-adic-numbers-in-physics
<p>As far as I know, in modern physics we assume that the underlying field of work is the field of real numbers (or complex numbers). Imagine one second that we make a crazy assumption and suggest that the fundamental equations of physics can be expressed with $p$-adic numbers. What could be really rewritten formally? Does it make sense in the physical world to work over $p$-adic numbers. Is there such attempt in the history of mathematical physics ? What is the most convincing justification (in physics) that we need to work over the field of real numbers ? </p>
[ { "answer_id": 259156, "score": 29, "text": "<p>For an overview of applications of p-adic numbers in physics I would refer to the <A HREF=\"https://en.wikipedia.org/wiki/P-adic_quantum_mechanics\">Wikipedia</A> and <A HREF=\"http://physics.stackexchange.com/questions/15252/p-adic-quantum-mechanic\">Physics.stackexchange</A> links, and to this <A HREF=\"https://ncatlab.org/nlab/show/p-adic+physics\">nLab entry.</A> Regarding the second question \"<em>What is the most convincing justification in physics that we need to work over the field of real or complex numbers</em>\" I would like to quote Freeman Dyson in <A HREF=\"http://www.ams.org/notices/200902/rtx090200212p.pdf\">Birds and Frogs:</A></p>\n\n<blockquote>\n <p>When I look at the history of mathematics, I see a succession of\n illogical jumps, improbable coincidences, jokes of nature. One of the\n most profound jokes of nature is the square root of minus one that the\n physicist Erwin Schrödinger put into his wave equation when he\n invented wave mechanics in 1926. Starting from wave optics as a model,\n he wrote down a differential equation for a mechanical particle, but\n the equation made no sense. The equation looked like the equation of\n conduction of heat in a continuous medium. Heat conduction has no\n visible relevance to particle mechanics. Schrödinger’s idea seemed to\n be going nowhere. But then came the surprise. Schrödinger put the\n square root of minus one into the equation, and suddenly it made\n sense. Suddenly it became a wave equation instead of a heat conduction\n equation. And Schrödinger found to his delight that the equation has\n solutions corresponding to the quantized orbits in the Bohr model of\n the atom.</p>\n \n <p>It turns out that the Schrödinger equation describes correctly\n everything we know about the behavior of atoms. It is the basis of all\n of chemistry and most of physics. <strong>And that square root of minus one\n means that nature works with complex numbers and not with real\n numbers.</strong></p>\n \n <p>All through the nineteenth century, mathematicians from Abel to\n Riemann and Weierstrass had been creating a magnificent theory of\n functions of complex variables. They had discovered that the theory of\n functions became far deeper and more powerful when it was extended\n from real to complex numbers. But they always thought of complex\n numbers as an artificial construction, invented by human\n mathematicians as a useful and elegant abstraction from real life. It\n never entered their heads that this artificial number system that they\n had invented was in fact the ground on which atoms move. They never\n imagined that nature had got there first.</p>\n</blockquote>\n" }, { "answer_id": 259160, "score": 24, "text": "<p>Since the <a href=\"https://physics.stackexchange.com/questions/15252/p-adic-quantum-mechanic\">question on physics.stackexchange</a> that was referred to in the comments above has been closed, let me essentially repeat my answer here together with some updates.</p>\n\n<hr>\n\n<p>I think there are mainly two reasons for motivating the introduction of <span class=\"math-container\">$p$</span>-adic models in physics.</p>\n\n<ol>\n<li>They could exist in nature.</li>\n<li>They provide insightful toy models for physical phenomena.</li>\n</ol>\n\n<p>There is a vast physical literature on the subject which cite Reason 1 as justification, this is sometimes called the Vladimirov Hypothesis. Namely, we do not know the texture of spacetime at the Planck scale therefore it is possible that it might look more like <span class=\"math-container\">$\\mathbb{Q}_p^d$</span> than <span class=\"math-container\">$\\mathbb{R}^d$</span>.\nIt is a seductive idea, but there is no evidence for it. Moreover, this would beg the question of why would Nature choose a particular prime number. A perhaps better hypothesis is the Adelic one advocated by Manin, i.e., all primes should occur democratically.\nIn any case, this is quite speculative at present.</p>\n\n<p>The way I see it,\n<span class=\"math-container\">$p$</span>-adic numbers are useful in physics mainly because of Reason 2.\nBy exaggerating only a little, one could argue that <span class=\"math-container\">$p$</span>-adic toy models is what guided Kenneth Wilson when making his great discoveries in the theory of the renormalization group\nwhich entirely revolutionized physics in the early seventies. </p>\n\n<p>When studying complex multiscale phenomena it is often important to decompose functions into time-frequency atoms which live on a tree, e.g., when using a wavelet decomposition.\nUnfortunately for most questions of interest the metric which governs how these atoms interact with each other is not the natural (from the tree point of view) ultrametric distance, but the Euclidean metric of the underlying continuum. Hierarchical models in physics amount to changing the model so it is the ultrametric distance which defines atomic interactions. The same idea also appears in mathematics where such toy models are often called \"dyadic models\". See <a href=\"https://terrytao.wordpress.com/2007/07/27/dyadic-models/\" rel=\"nofollow noreferrer\">this wonderful post</a> by Tao for a nice discussion of this circle of ideas. Given a problem in Euclidean space, there are lots of ways of setting up a simplified hierarchical model for it.\nThe <span class=\"math-container\">$p$</span>-adics, in some sense provide the most canonical, structured and principled way of doing this.</p>\n\n<p>To go back to Wilson and the RG, he also used a hierarchical model\nand gave it yet another name \"the approximate recursion\". The importance of this toy model in Wilson's path to discovery is clear from his quote:</p>\n\n<blockquote>\n <p>\"Then, at Michael's urging, I work out what happens near four dimensions for the approximate recursion formula, and find that d-4 acts as a small parameter. Knowing this it is then trivial, given my field theoretic training, to construct the beginning of the epsilon expansion for critical exponents.\"</p>\n</blockquote>\n\n<p>which can be found <a href=\"http://authors.library.caltech.edu/5456/1/hrst.mit.edu/hrs/renormalization/Wilson/Wilson2.htm\" rel=\"nofollow noreferrer\">here</a> (about one third from the bottom of the page).</p>\n\n<p>For example, the use of <span class=\"math-container\">$p$</span>-adics gives a toy model for rigorously understanding conformal field theories. This is explained in my article <a href=\"https://arxiv.org/abs/1511.03180\" rel=\"nofollow noreferrer\">\"Towards three-dimensional conformal probability\"</a> (<a href=\"http://people.virginia.edu/~aa4cr/TalkUMDfinal.pdf\" rel=\"nofollow noreferrer\">these slides</a> are perhaps easier to understand since they have lots of pictures). \nMore recently, the following articles explored similar issues:</p>\n\n<ol>\n<li><a href=\"https://arxiv.org/abs/1605.01061\" rel=\"nofollow noreferrer\">\"<span class=\"math-container\">$p$</span>-adic AdS/CFT\"</a> by Gubser, Knaute, Parikh, Samberg and Witaszczyk,</li>\n<li><a href=\"https://arxiv.org/abs/1605.07639\" rel=\"nofollow noreferrer\">\"Tensor networks, <span class=\"math-container\">$p$</span>-adic fields, and algebraic curves: arithmetic and the <span class=\"math-container\">${\\rm AdS}_3$</span>/<span class=\"math-container\">${\\rm CFT}_2$</span> correspondence\"</a> by Heydeman, Marcolli, Saberi and Stoica,</li>\n<li><a href=\"https://arxiv.org/abs/1612.09580\" rel=\"nofollow noreferrer\">\"Edge length dynamics on graphs with applications to <span class=\"math-container\">$p$</span>-adic AdS/CFT\"</a> by Gubser, Heydeman, Jepsen, Marcolli, Parikh, Saberi, Stoica and Trundy. </li>\n</ol>\n\n<p>As another example, in the article <a href=\"https://link.springer.com/article/10.1007/s00220-019-03665-4\" rel=\"nofollow noreferrer\">\"A second-quantized Kolmogorov-Chentsov theorem via the operator product expansion\"</a> I discovered a general result for making sense of products of random Schwartz distributions starting from Wilson's operator product expansion. The statement of the theorem and the method of proof was done over <span class=\"math-container\">$\\mathbb{Q}_p$</span>\nfirst.</p>\n\n<p>Another remark is that exotic numbers like <span class=\"math-container\">$p$</span>-adics can appear in two different conceptual roles. Suppose one is interested in a \"scalar field\" <span class=\"math-container\">$\\phi:X\\rightarrow Y$</span> where <span class=\"math-container\">$X$</span> is space-time. The toy models I mentioned make <span class=\"math-container\">$X$</span> <span class=\"math-container\">$p$</span>-adic but keep <span class=\"math-container\">$Y$</span> real or complex. As in Dyson's quote from Carlo's answer, things should still be Archimedean-valued. This kind of distinction is well known to researchers in the area called the Langlands Program which splits in two separate arenas depending on whether one's favorite <span class=\"math-container\">$L$</span> or zeta functions take values in <span class=\"math-container\">$\\mathbb{C}$</span> or <span class=\"math-container\">$\\mathbb{Q}_p$</span>.\nThere are works in <span class=\"math-container\">$p$</span>-adic physics (e.g. by Khrennikov) where <span class=\"math-container\">$Y$</span> is taken <span class=\"math-container\">$p$</span>-adic, but I am not familiar with the motivation behind this.</p>\n\n<p>Finally, and in relation to Laie's comment above, one possible use of <span class=\"math-container\">$p$</span>-adics is to try to exploit an Adelic product formula in order to compute a real quantity of interest as the inverse of the product of similar quantitites over <span class=\"math-container\">$\\mathbb{Q}_p$</span>. This was observed for tree level Veneziano amplitudes and was one of the first motivations for developing <span class=\"math-container\">$p$</span>-adic physics. However, these adelic formulas break down for higher loop amplitudes. See <a href=\"http://www.sciencedirect.com/science/article/pii/037015739390043D\" rel=\"nofollow noreferrer\">this Physics Report article</a> for more on this story.</p>\n" }, { "answer_id": 259156, "score": 29, "text": "<p>For an overview of applications of p-adic numbers in physics I would refer to the <A HREF=\"https://en.wikipedia.org/wiki/P-adic_quantum_mechanics\">Wikipedia</A> and <A HREF=\"http://physics.stackexchange.com/questions/15252/p-adic-quantum-mechanic\">Physics.stackexchange</A> links, and to this <A HREF=\"https://ncatlab.org/nlab/show/p-adic+physics\">nLab entry.</A> Regarding the second question \"<em>What is the most convincing justification in physics that we need to work over the field of real or complex numbers</em>\" I would like to quote Freeman Dyson in <A HREF=\"http://www.ams.org/notices/200902/rtx090200212p.pdf\">Birds and Frogs:</A></p>\n\n<blockquote>\n <p>When I look at the history of mathematics, I see a succession of\n illogical jumps, improbable coincidences, jokes of nature. One of the\n most profound jokes of nature is the square root of minus one that the\n physicist Erwin Schrödinger put into his wave equation when he\n invented wave mechanics in 1926. Starting from wave optics as a model,\n he wrote down a differential equation for a mechanical particle, but\n the equation made no sense. The equation looked like the equation of\n conduction of heat in a continuous medium. Heat conduction has no\n visible relevance to particle mechanics. Schrödinger’s idea seemed to\n be going nowhere. But then came the surprise. Schrödinger put the\n square root of minus one into the equation, and suddenly it made\n sense. Suddenly it became a wave equation instead of a heat conduction\n equation. And Schrödinger found to his delight that the equation has\n solutions corresponding to the quantized orbits in the Bohr model of\n the atom.</p>\n \n <p>It turns out that the Schrödinger equation describes correctly\n everything we know about the behavior of atoms. It is the basis of all\n of chemistry and most of physics. <strong>And that square root of minus one\n means that nature works with complex numbers and not with real\n numbers.</strong></p>\n \n <p>All through the nineteenth century, mathematicians from Abel to\n Riemann and Weierstrass had been creating a magnificent theory of\n functions of complex variables. They had discovered that the theory of\n functions became far deeper and more powerful when it was extended\n from real to complex numbers. But they always thought of complex\n numbers as an artificial construction, invented by human\n mathematicians as a useful and elegant abstraction from real life. It\n never entered their heads that this artificial number system that they\n had invented was in fact the ground on which atoms move. They never\n imagined that nature had got there first.</p>\n</blockquote>\n" }, { "answer_id": 259160, "score": 24, "text": "<p>Since the <a href=\"https://physics.stackexchange.com/questions/15252/p-adic-quantum-mechanic\">question on physics.stackexchange</a> that was referred to in the comments above has been closed, let me essentially repeat my answer here together with some updates.</p>\n\n<hr>\n\n<p>I think there are mainly two reasons for motivating the introduction of <span class=\"math-container\">$p$</span>-adic models in physics.</p>\n\n<ol>\n<li>They could exist in nature.</li>\n<li>They provide insightful toy models for physical phenomena.</li>\n</ol>\n\n<p>There is a vast physical literature on the subject which cite Reason 1 as justification, this is sometimes called the Vladimirov Hypothesis. Namely, we do not know the texture of spacetime at the Planck scale therefore it is possible that it might look more like <span class=\"math-container\">$\\mathbb{Q}_p^d$</span> than <span class=\"math-container\">$\\mathbb{R}^d$</span>.\nIt is a seductive idea, but there is no evidence for it. Moreover, this would beg the question of why would Nature choose a particular prime number. A perhaps better hypothesis is the Adelic one advocated by Manin, i.e., all primes should occur democratically.\nIn any case, this is quite speculative at present.</p>\n\n<p>The way I see it,\n<span class=\"math-container\">$p$</span>-adic numbers are useful in physics mainly because of Reason 2.\nBy exaggerating only a little, one could argue that <span class=\"math-container\">$p$</span>-adic toy models is what guided Kenneth Wilson when making his great discoveries in the theory of the renormalization group\nwhich entirely revolutionized physics in the early seventies. </p>\n\n<p>When studying complex multiscale phenomena it is often important to decompose functions into time-frequency atoms which live on a tree, e.g., when using a wavelet decomposition.\nUnfortunately for most questions of interest the metric which governs how these atoms interact with each other is not the natural (from the tree point of view) ultrametric distance, but the Euclidean metric of the underlying continuum. Hierarchical models in physics amount to changing the model so it is the ultrametric distance which defines atomic interactions. The same idea also appears in mathematics where such toy models are often called \"dyadic models\". See <a href=\"https://terrytao.wordpress.com/2007/07/27/dyadic-models/\" rel=\"nofollow noreferrer\">this wonderful post</a> by Tao for a nice discussion of this circle of ideas. Given a problem in Euclidean space, there are lots of ways of setting up a simplified hierarchical model for it.\nThe <span class=\"math-container\">$p$</span>-adics, in some sense provide the most canonical, structured and principled way of doing this.</p>\n\n<p>To go back to Wilson and the RG, he also used a hierarchical model\nand gave it yet another name \"the approximate recursion\". The importance of this toy model in Wilson's path to discovery is clear from his quote:</p>\n\n<blockquote>\n <p>\"Then, at Michael's urging, I work out what happens near four dimensions for the approximate recursion formula, and find that d-4 acts as a small parameter. Knowing this it is then trivial, given my field theoretic training, to construct the beginning of the epsilon expansion for critical exponents.\"</p>\n</blockquote>\n\n<p>which can be found <a href=\"http://authors.library.caltech.edu/5456/1/hrst.mit.edu/hrs/renormalization/Wilson/Wilson2.htm\" rel=\"nofollow noreferrer\">here</a> (about one third from the bottom of the page).</p>\n\n<p>For example, the use of <span class=\"math-container\">$p$</span>-adics gives a toy model for rigorously understanding conformal field theories. This is explained in my article <a href=\"https://arxiv.org/abs/1511.03180\" rel=\"nofollow noreferrer\">\"Towards three-dimensional conformal probability\"</a> (<a href=\"http://people.virginia.edu/~aa4cr/TalkUMDfinal.pdf\" rel=\"nofollow noreferrer\">these slides</a> are perhaps easier to understand since they have lots of pictures). \nMore recently, the following articles explored similar issues:</p>\n\n<ol>\n<li><a href=\"https://arxiv.org/abs/1605.01061\" rel=\"nofollow noreferrer\">\"<span class=\"math-container\">$p$</span>-adic AdS/CFT\"</a> by Gubser, Knaute, Parikh, Samberg and Witaszczyk,</li>\n<li><a href=\"https://arxiv.org/abs/1605.07639\" rel=\"nofollow noreferrer\">\"Tensor networks, <span class=\"math-container\">$p$</span>-adic fields, and algebraic curves: arithmetic and the <span class=\"math-container\">${\\rm AdS}_3$</span>/<span class=\"math-container\">${\\rm CFT}_2$</span> correspondence\"</a> by Heydeman, Marcolli, Saberi and Stoica,</li>\n<li><a href=\"https://arxiv.org/abs/1612.09580\" rel=\"nofollow noreferrer\">\"Edge length dynamics on graphs with applications to <span class=\"math-container\">$p$</span>-adic AdS/CFT\"</a> by Gubser, Heydeman, Jepsen, Marcolli, Parikh, Saberi, Stoica and Trundy. </li>\n</ol>\n\n<p>As another example, in the article <a href=\"https://link.springer.com/article/10.1007/s00220-019-03665-4\" rel=\"nofollow noreferrer\">\"A second-quantized Kolmogorov-Chentsov theorem via the operator product expansion\"</a> I discovered a general result for making sense of products of random Schwartz distributions starting from Wilson's operator product expansion. The statement of the theorem and the method of proof was done over <span class=\"math-container\">$\\mathbb{Q}_p$</span>\nfirst.</p>\n\n<p>Another remark is that exotic numbers like <span class=\"math-container\">$p$</span>-adics can appear in two different conceptual roles. Suppose one is interested in a \"scalar field\" <span class=\"math-container\">$\\phi:X\\rightarrow Y$</span> where <span class=\"math-container\">$X$</span> is space-time. The toy models I mentioned make <span class=\"math-container\">$X$</span> <span class=\"math-container\">$p$</span>-adic but keep <span class=\"math-container\">$Y$</span> real or complex. As in Dyson's quote from Carlo's answer, things should still be Archimedean-valued. This kind of distinction is well known to researchers in the area called the Langlands Program which splits in two separate arenas depending on whether one's favorite <span class=\"math-container\">$L$</span> or zeta functions take values in <span class=\"math-container\">$\\mathbb{C}$</span> or <span class=\"math-container\">$\\mathbb{Q}_p$</span>.\nThere are works in <span class=\"math-container\">$p$</span>-adic physics (e.g. by Khrennikov) where <span class=\"math-container\">$Y$</span> is taken <span class=\"math-container\">$p$</span>-adic, but I am not familiar with the motivation behind this.</p>\n\n<p>Finally, and in relation to Laie's comment above, one possible use of <span class=\"math-container\">$p$</span>-adics is to try to exploit an Adelic product formula in order to compute a real quantity of interest as the inverse of the product of similar quantitites over <span class=\"math-container\">$\\mathbb{Q}_p$</span>. This was observed for tree level Veneziano amplitudes and was one of the first motivations for developing <span class=\"math-container\">$p$</span>-adic physics. However, these adelic formulas break down for higher loop amplitudes. See <a href=\"http://www.sciencedirect.com/science/article/pii/037015739390043D\" rel=\"nofollow noreferrer\">this Physics Report article</a> for more on this story.</p>\n" }, { "answer_id": 259156, "score": 29, "text": "<p>For an overview of applications of p-adic numbers in physics I would refer to the <A HREF=\"https://en.wikipedia.org/wiki/P-adic_quantum_mechanics\">Wikipedia</A> and <A HREF=\"http://physics.stackexchange.com/questions/15252/p-adic-quantum-mechanic\">Physics.stackexchange</A> links, and to this <A HREF=\"https://ncatlab.org/nlab/show/p-adic+physics\">nLab entry.</A> Regarding the second question \"<em>What is the most convincing justification in physics that we need to work over the field of real or complex numbers</em>\" I would like to quote Freeman Dyson in <A HREF=\"http://www.ams.org/notices/200902/rtx090200212p.pdf\">Birds and Frogs:</A></p>\n\n<blockquote>\n <p>When I look at the history of mathematics, I see a succession of\n illogical jumps, improbable coincidences, jokes of nature. One of the\n most profound jokes of nature is the square root of minus one that the\n physicist Erwin Schrödinger put into his wave equation when he\n invented wave mechanics in 1926. Starting from wave optics as a model,\n he wrote down a differential equation for a mechanical particle, but\n the equation made no sense. The equation looked like the equation of\n conduction of heat in a continuous medium. Heat conduction has no\n visible relevance to particle mechanics. Schrödinger’s idea seemed to\n be going nowhere. But then came the surprise. Schrödinger put the\n square root of minus one into the equation, and suddenly it made\n sense. Suddenly it became a wave equation instead of a heat conduction\n equation. And Schrödinger found to his delight that the equation has\n solutions corresponding to the quantized orbits in the Bohr model of\n the atom.</p>\n \n <p>It turns out that the Schrödinger equation describes correctly\n everything we know about the behavior of atoms. It is the basis of all\n of chemistry and most of physics. <strong>And that square root of minus one\n means that nature works with complex numbers and not with real\n numbers.</strong></p>\n \n <p>All through the nineteenth century, mathematicians from Abel to\n Riemann and Weierstrass had been creating a magnificent theory of\n functions of complex variables. They had discovered that the theory of\n functions became far deeper and more powerful when it was extended\n from real to complex numbers. But they always thought of complex\n numbers as an artificial construction, invented by human\n mathematicians as a useful and elegant abstraction from real life. It\n never entered their heads that this artificial number system that they\n had invented was in fact the ground on which atoms move. They never\n imagined that nature had got there first.</p>\n</blockquote>\n" }, { "answer_id": 259160, "score": 24, "text": "<p>Since the <a href=\"https://physics.stackexchange.com/questions/15252/p-adic-quantum-mechanic\">question on physics.stackexchange</a> that was referred to in the comments above has been closed, let me essentially repeat my answer here together with some updates.</p>\n\n<hr>\n\n<p>I think there are mainly two reasons for motivating the introduction of <span class=\"math-container\">$p$</span>-adic models in physics.</p>\n\n<ol>\n<li>They could exist in nature.</li>\n<li>They provide insightful toy models for physical phenomena.</li>\n</ol>\n\n<p>There is a vast physical literature on the subject which cite Reason 1 as justification, this is sometimes called the Vladimirov Hypothesis. Namely, we do not know the texture of spacetime at the Planck scale therefore it is possible that it might look more like <span class=\"math-container\">$\\mathbb{Q}_p^d$</span> than <span class=\"math-container\">$\\mathbb{R}^d$</span>.\nIt is a seductive idea, but there is no evidence for it. Moreover, this would beg the question of why would Nature choose a particular prime number. A perhaps better hypothesis is the Adelic one advocated by Manin, i.e., all primes should occur democratically.\nIn any case, this is quite speculative at present.</p>\n\n<p>The way I see it,\n<span class=\"math-container\">$p$</span>-adic numbers are useful in physics mainly because of Reason 2.\nBy exaggerating only a little, one could argue that <span class=\"math-container\">$p$</span>-adic toy models is what guided Kenneth Wilson when making his great discoveries in the theory of the renormalization group\nwhich entirely revolutionized physics in the early seventies. </p>\n\n<p>When studying complex multiscale phenomena it is often important to decompose functions into time-frequency atoms which live on a tree, e.g., when using a wavelet decomposition.\nUnfortunately for most questions of interest the metric which governs how these atoms interact with each other is not the natural (from the tree point of view) ultrametric distance, but the Euclidean metric of the underlying continuum. Hierarchical models in physics amount to changing the model so it is the ultrametric distance which defines atomic interactions. The same idea also appears in mathematics where such toy models are often called \"dyadic models\". See <a href=\"https://terrytao.wordpress.com/2007/07/27/dyadic-models/\" rel=\"nofollow noreferrer\">this wonderful post</a> by Tao for a nice discussion of this circle of ideas. Given a problem in Euclidean space, there are lots of ways of setting up a simplified hierarchical model for it.\nThe <span class=\"math-container\">$p$</span>-adics, in some sense provide the most canonical, structured and principled way of doing this.</p>\n\n<p>To go back to Wilson and the RG, he also used a hierarchical model\nand gave it yet another name \"the approximate recursion\". The importance of this toy model in Wilson's path to discovery is clear from his quote:</p>\n\n<blockquote>\n <p>\"Then, at Michael's urging, I work out what happens near four dimensions for the approximate recursion formula, and find that d-4 acts as a small parameter. Knowing this it is then trivial, given my field theoretic training, to construct the beginning of the epsilon expansion for critical exponents.\"</p>\n</blockquote>\n\n<p>which can be found <a href=\"http://authors.library.caltech.edu/5456/1/hrst.mit.edu/hrs/renormalization/Wilson/Wilson2.htm\" rel=\"nofollow noreferrer\">here</a> (about one third from the bottom of the page).</p>\n\n<p>For example, the use of <span class=\"math-container\">$p$</span>-adics gives a toy model for rigorously understanding conformal field theories. This is explained in my article <a href=\"https://arxiv.org/abs/1511.03180\" rel=\"nofollow noreferrer\">\"Towards three-dimensional conformal probability\"</a> (<a href=\"http://people.virginia.edu/~aa4cr/TalkUMDfinal.pdf\" rel=\"nofollow noreferrer\">these slides</a> are perhaps easier to understand since they have lots of pictures). \nMore recently, the following articles explored similar issues:</p>\n\n<ol>\n<li><a href=\"https://arxiv.org/abs/1605.01061\" rel=\"nofollow noreferrer\">\"<span class=\"math-container\">$p$</span>-adic AdS/CFT\"</a> by Gubser, Knaute, Parikh, Samberg and Witaszczyk,</li>\n<li><a href=\"https://arxiv.org/abs/1605.07639\" rel=\"nofollow noreferrer\">\"Tensor networks, <span class=\"math-container\">$p$</span>-adic fields, and algebraic curves: arithmetic and the <span class=\"math-container\">${\\rm AdS}_3$</span>/<span class=\"math-container\">${\\rm CFT}_2$</span> correspondence\"</a> by Heydeman, Marcolli, Saberi and Stoica,</li>\n<li><a href=\"https://arxiv.org/abs/1612.09580\" rel=\"nofollow noreferrer\">\"Edge length dynamics on graphs with applications to <span class=\"math-container\">$p$</span>-adic AdS/CFT\"</a> by Gubser, Heydeman, Jepsen, Marcolli, Parikh, Saberi, Stoica and Trundy. </li>\n</ol>\n\n<p>As another example, in the article <a href=\"https://link.springer.com/article/10.1007/s00220-019-03665-4\" rel=\"nofollow noreferrer\">\"A second-quantized Kolmogorov-Chentsov theorem via the operator product expansion\"</a> I discovered a general result for making sense of products of random Schwartz distributions starting from Wilson's operator product expansion. The statement of the theorem and the method of proof was done over <span class=\"math-container\">$\\mathbb{Q}_p$</span>\nfirst.</p>\n\n<p>Another remark is that exotic numbers like <span class=\"math-container\">$p$</span>-adics can appear in two different conceptual roles. Suppose one is interested in a \"scalar field\" <span class=\"math-container\">$\\phi:X\\rightarrow Y$</span> where <span class=\"math-container\">$X$</span> is space-time. The toy models I mentioned make <span class=\"math-container\">$X$</span> <span class=\"math-container\">$p$</span>-adic but keep <span class=\"math-container\">$Y$</span> real or complex. As in Dyson's quote from Carlo's answer, things should still be Archimedean-valued. This kind of distinction is well known to researchers in the area called the Langlands Program which splits in two separate arenas depending on whether one's favorite <span class=\"math-container\">$L$</span> or zeta functions take values in <span class=\"math-container\">$\\mathbb{C}$</span> or <span class=\"math-container\">$\\mathbb{Q}_p$</span>.\nThere are works in <span class=\"math-container\">$p$</span>-adic physics (e.g. by Khrennikov) where <span class=\"math-container\">$Y$</span> is taken <span class=\"math-container\">$p$</span>-adic, but I am not familiar with the motivation behind this.</p>\n\n<p>Finally, and in relation to Laie's comment above, one possible use of <span class=\"math-container\">$p$</span>-adics is to try to exploit an Adelic product formula in order to compute a real quantity of interest as the inverse of the product of similar quantitites over <span class=\"math-container\">$\\mathbb{Q}_p$</span>. This was observed for tree level Veneziano amplitudes and was one of the first motivations for developing <span class=\"math-container\">$p$</span>-adic physics. However, these adelic formulas break down for higher loop amplitudes. See <a href=\"http://www.sciencedirect.com/science/article/pii/037015739390043D\" rel=\"nofollow noreferrer\">this Physics Report article</a> for more on this story.</p>\n" } ]
Question: $p$-adic numbers in physics Answer 1: <p>For an overview of applications of p-adic numbers in physics I would refer to the <A HREF="https://en.wikipedia.org/wiki/P-adic_quantum_mechanics">Wikipedia</A> and <A HREF="http://physics.stackexchange.com/questions/15252/p-adic-quantum-mechanic">Physics.stackexchange</A> links, and to this <A HREF="https://ncatlab.org/nlab/show/p-adic+physics">nLab entry.</A> Regarding the second question "<em>What is the most convincing justification in physics that we need to work over the field of real or complex numbers</em>" I would like to quote Freeman Dyson in <A HREF="http://www.ams.org/notices/200902/rtx090200212p.pdf">Birds and Frogs:</A></p> <blockquote> <p>When I look at the history of mathematics, I see a succession of illogical jumps, improbable coincidences, jokes of nature. One of the most profound jokes of nature is the square root of minus one that the physicist Erwin Schrödinger put into his wave equation when he invented wave mechanics in 1926. Starting from wave optics as a model, he wrote down a differential equation for a mechanical particle, but the equation made no sense. The equation looked like the equation of conduction of heat in a continuous medium. Heat conduction has no visible relevance to particle mechanics. Schrödinger’s idea seemed to be going nowhere. But then came the surprise. Schrödinger put the square root of minus one into the equation, and suddenly it made sense. Suddenly it became a wave equation instead of a heat conduction equation. And Schrödinger found to his delight that the equation has solutions corresponding to the quantized orbits in the Bohr model of the atom.</p> <p>It turns out that the Schrödinger equation describes correctly everything we know about the behavior of atoms. It is the basis of all of chemistry and most of physics. <strong>And that square root of minus one means that nature works with complex numbers and not with real numbers.</strong></p> <p>All through the nineteenth century, mathematicians from Abel to Riemann and Weierstrass had been creating a magnificent theory of functions of complex variables. They had discovered that the theory of functions became far deeper and more powerful when it was extended from real to complex numbers. But they always thought of complex numbers as an artificial construction, invented by human mathematicians as a useful and elegant abstraction from real life. It never entered their heads that this artificial number system that they had invented was in fact the ground on which atoms move. They never imagined that nature had got there first.</p> </blockquote> Answer 2: <p>Since the <a href="https://physics.stackexchange.com/questions/15252/p-adic-quantum-mechanic">question on physics.stackexchange</a> that was referred to in the comments above has been closed, let me essentially repeat my answer here together with some updates.</p> <hr> <p>I think there are mainly two reasons for motivating the introduction of <span class="math-container">$p$</span>-adic models in physics.</p> <ol> <li>They could exist in nature.</li> <li>They provide insightful toy models for physical phenomena.</li> </ol> <p>There is a vast physical literature on the subject which cite Reason 1 as justification, this is sometimes called the Vladimirov Hypothesis. Namely, we do not know the texture of spacetime at the Planck scale therefore it is possible that it might look more like <span class="math-container">$\mathbb{Q}_p^d$</span> than <span class="math-container">$\mathbb{R}^d$</span>. It is a seductive idea, but there is no evidence for it. Moreover, this would beg the question of why would Nature choose a particular prime number. A perhaps better hypothesis is the Adelic one advocated by Manin, i.e., all primes should occur democratically. In any case, this is quite speculative at present.</p> <p>The way I see it, <span class="math-container">$p$</span>-adic numbers are useful in physics mainly because of Reason 2. By exaggerating only a little, one could argue that <span class="math-container">$p$</span>-adic toy models is what guided Kenneth Wilson when making his great discoveries in the theory of the renormalization group which entirely revolutionized physics in the early seventies. </p> <p>When studying complex multiscale phenomena it is often important to decompose functions into time-frequency atoms which live on a tree, e.g., when using a wavelet decomposition. Unfortunately for most questions of interest the metric which governs how these atoms interact with each other is not the natural (from the tree point of view) ultrametric distance, but the Euclidean metric of the underlying continuum. Hierarchical models in physics amount to changing the model so it is the ultrametric distance which defines atomic interactions. The same idea also appears in mathematics where such toy models are often called "dyadic models". See <a href="https://terrytao.wordpress.com/2007/07/27/dyadic-models/" rel="nofollow noreferrer">this wonderful post</a> by Tao for a nice discussion of this circle of ideas. Given a problem in Euclidean space, there are lots of ways of setting up a simplified hierarchical model for it. The <span class="math-container">$p$</span>-adics, in some sense provide the most canonical, structured and principled way of doing this.</p> <p>To go back to Wilson and the RG, he also used a hierarchical model and gave it yet another name "the approximate recursion". The importance of this toy model in Wilson's path to discovery is clear from his quote:</p> <blockquote> <p>"Then, at Michael's urging, I work out what happens near four dimensions for the approximate recursion formula, and find that d-4 acts as a small parameter. Knowing this it is then trivial, given my field theoretic training, to construct the beginning of the epsilon expansion for critical exponents."</p> </blockquote> <p>which can be found <a href="http://authors.library.caltech.edu/5456/1/hrst.mit.edu/hrs/renormalization/Wilson/Wilson2.htm" rel="nofollow noreferrer">here</a> (about one third from the bottom of the page).</p> <p>For example, the use of <span class="math-container">$p$</span>-adics gives a toy model for rigorously understanding conformal field theories. This is explained in my article <a href="https://arxiv.org/abs/1511.03180" rel="nofollow noreferrer">"Towards three-dimensional conformal probability"</a> (<a href="http://people.virginia.edu/~aa4cr/TalkUMDfinal.pdf" rel="nofollow noreferrer">these slides</a> are perhaps easier to understand since they have lots of pictures). More recently, the following articles explored similar issues:</p> <ol> <li><a href="https://arxiv.org/abs/1605.01061" rel="nofollow noreferrer">"<span class="math-container">$p$</span>-adic AdS/CFT"</a> by Gubser, Knaute, Parikh, Samberg and Witaszczyk,</li> <li><a href="https://arxiv.org/abs/1605.07639" rel="nofollow noreferrer">"Tensor networks, <span class="math-container">$p$</span>-adic fields, and algebraic curves: arithmetic and the <span class="math-container">${\rm AdS}_3$</span>/<span class="math-container">${\rm CFT}_2$</span> correspondence"</a> by Heydeman, Marcolli, Saberi and Stoica,</li> <li><a href="https://arxiv.org/abs/1612.09580" rel="nofollow noreferrer">"Edge length dynamics on graphs with applications to <span class="math-container">$p$</span>-adic AdS/CFT"</a> by Gubser, Heydeman, Jepsen, Marcolli, Parikh, Saberi, Stoica and Trundy. </li> </ol> <p>As another example, in the article <a href="https://link.springer.com/article/10.1007/s00220-019-03665-4" rel="nofollow noreferrer">"A second-quantized Kolmogorov-Chentsov theorem via the operator product expansion"</a> I discovered a general result for making sense of products of random Schwartz distributions starting from Wilson's operator product expansion. The statement of the theorem and the method of proof was done over <span class="math-container">$\mathbb{Q}_p$</span> first.</p> <p>Another remark is that exotic numbers like <span class="math-container">$p$</span>-adics can appear in two different conceptual roles. Suppose one is interested in a "scalar field" <span class="math-container">$\phi:X\rightarrow Y$</span> where <span class="math-container">$X$</span> is space-time. The toy models I mentioned make <span class="math-container">$X$</span> <span class="math-container">$p$</span>-adic but keep <span class="math-container">$Y$</span> real or complex. As in Dyson's quote from Carlo's answer, things should still be Archimedean-valued. This kind of distinction is well known to researchers in the area called the Langlands Program which splits in two separate arenas depending on whether one's favorite <span class="math-container">$L$</span> or zeta functions take values in <span class="math-container">$\mathbb{C}$</span> or <span class="math-container">$\mathbb{Q}_p$</span>. There are works in <span class="math-container">$p$</span>-adic physics (e.g. by Khrennikov) where <span class="math-container">$Y$</span> is taken <span class="math-container">$p$</span>-adic, but I am not familiar with the motivation behind this.</p> <p>Finally, and in relation to Laie's comment above, one possible use of <span class="math-container">$p$</span>-adics is to try to exploit an Adelic product formula in order to compute a real quantity of interest as the inverse of the product of similar quantitites over <span class="math-container">$\mathbb{Q}_p$</span>. This was observed for tree level Veneziano amplitudes and was one of the first motivations for developing <span class="math-container">$p$</span>-adic physics. However, these adelic formulas break down for higher loop amplitudes. See <a href="http://www.sciencedirect.com/science/article/pii/037015739390043D" rel="nofollow noreferrer">this Physics Report article</a> for more on this story.</p> Answer 3: <p>For an overview of applications of p-adic numbers in physics I would refer to the <A HREF="https://en.wikipedia.org/wiki/P-adic_quantum_mechanics">Wikipedia</A> and <A HREF="http://physics.stackexchange.com/questions/15252/p-adic-quantum-mechanic">Physics.stackexchange</A> links, and to this <A HREF="https://ncatlab.org/nlab/show/p-adic+physics">nLab entry.</A> Regarding the second question "<em>What is the most convincing justification in physics that we need to work over the field of real or complex numbers</em>" I would like to quote Freeman Dyson in <A HREF="http://www.ams.org/notices/200902/rtx090200212p.pdf">Birds and Frogs:</A></p> <blockquote> <p>When I look at the history of mathematics, I see a succession of illogical jumps, improbable coincidences, jokes of nature. One of the most profound jokes of nature is the square root of minus one that the physicist Erwin Schrödinger put into his wave equation when he invented wave mechanics in 1926. Starting from wave optics as a model, he wrote down a differential equation for a mechanical particle, but the equation made no sense. The equation looked like the equation of conduction of heat in a continuous medium. Heat conduction has no visible relevance to particle mechanics. Schrödinger’s idea seemed to be going nowhere. But then came the surprise. Schrödinger put the square root of minus one into the equation, and suddenly it made sense. Suddenly it became a wave equation instead of a heat conduction equation. And Schrödinger found to his delight that the equation has solutions corresponding to the quantized orbits in the Bohr model of the atom.</p> <p>It turns out that the Schrödinger equation describes correctly everything we know about the behavior of atoms. It is the basis of all of chemistry and most of physics. <strong>And that square root of minus one means that nature works with complex numbers and not with real numbers.</strong></p> <p>All through the nineteenth century, mathematicians from Abel to Riemann and Weierstrass had been creating a magnificent theory of functions of complex variables. They had discovered that the theory of functions became far deeper and more powerful when it was extended from real to complex numbers. But they always thought of complex numbers as an artificial construction, invented by human mathematicians as a useful and elegant abstraction from real life. It never entered their heads that this artificial number system that they had invented was in fact the ground on which atoms move. They never imagined that nature had got there first.</p> </blockquote> Answer 4: <p>Since the <a href="https://physics.stackexchange.com/questions/15252/p-adic-quantum-mechanic">question on physics.stackexchange</a> that was referred to in the comments above has been closed, let me essentially repeat my answer here together with some updates.</p> <hr> <p>I think there are mainly two reasons for motivating the introduction of <span class="math-container">$p$</span>-adic models in physics.</p> <ol> <li>They could exist in nature.</li> <li>They provide insightful toy models for physical phenomena.</li> </ol> <p>There is a vast physical literature on the subject which cite Reason 1 as justification, this is sometimes called the Vladimirov Hypothesis. Namely, we do not know the texture of spacetime at the Planck scale therefore it is possible that it might look more like <span class="math-container">$\mathbb{Q}_p^d$</span> than <span class="math-container">$\mathbb{R}^d$</span>. It is a seductive idea, but there is no evidence for it. Moreover, this would beg the question of why would Nature choose a particular prime number. A perhaps better hypothesis is the Adelic one advocated by Manin, i.e., all primes should occur democratically. In any case, this is quite speculative at present.</p> <p>The way I see it, <span class="math-container">$p$</span>-adic numbers are useful in physics mainly because of Reason 2. By exaggerating only a little, one could argue that <span class="math-container">$p$</span>-adic toy models is what guided Kenneth Wilson when making his great discoveries in the theory of the renormalization group which entirely revolutionized physics in the early seventies. </p> <p>When studying complex multiscale phenomena it is often important to decompose functions into time-frequency atoms which live on a tree, e.g., when using a wavelet decomposition. Unfortunately for most questions of interest the metric which governs how these atoms interact with each other is not the natural (from the tree point of view) ultrametric distance, but the Euclidean metric of the underlying continuum. Hierarchical models in physics amount to changing the model so it is the ultrametric distance which defines atomic interactions. The same idea also appears in mathematics where such toy models are often called "dyadic models". See <a href="https://terrytao.wordpress.com/2007/07/27/dyadic-models/" rel="nofollow noreferrer">this wonderful post</a> by Tao for a nice discussion of this circle of ideas. Given a problem in Euclidean space, there are lots of ways of setting up a simplified hierarchical model for it. The <span class="math-container">$p$</span>-adics, in some sense provide the most canonical, structured and principled way of doing this.</p> <p>To go back to Wilson and the RG, he also used a hierarchical model and gave it yet another name "the approximate recursion". The importance of this toy model in Wilson's path to discovery is clear from his quote:</p> <blockquote> <p>"Then, at Michael's urging, I work out what happens near four dimensions for the approximate recursion formula, and find that d-4 acts as a small parameter. Knowing this it is then trivial, given my field theoretic training, to construct the beginning of the epsilon expansion for critical exponents."</p> </blockquote> <p>which can be found <a href="http://authors.library.caltech.edu/5456/1/hrst.mit.edu/hrs/renormalization/Wilson/Wilson2.htm" rel="nofollow noreferrer">here</a> (about one third from the bottom of the page).</p> <p>For example, the use of <span class="math-container">$p$</span>-adics gives a toy model for rigorously understanding conformal field theories. This is explained in my article <a href="https://arxiv.org/abs/1511.03180" rel="nofollow noreferrer">"Towards three-dimensional conformal probability"</a> (<a href="http://people.virginia.edu/~aa4cr/TalkUMDfinal.pdf" rel="nofollow noreferrer">these slides</a> are perhaps easier to understand since they have lots of pictures). More recently, the following articles explored similar issues:</p> <ol> <li><a href="https://arxiv.org/abs/1605.01061" rel="nofollow noreferrer">"<span class="math-container">$p$</span>-adic AdS/CFT"</a> by Gubser, Knaute, Parikh, Samberg and Witaszczyk,</li> <li><a href="https://arxiv.org/abs/1605.07639" rel="nofollow noreferrer">"Tensor networks, <span class="math-container">$p$</span>-adic fields, and algebraic curves: arithmetic and the <span class="math-container">${\rm AdS}_3$</span>/<span class="math-container">${\rm CFT}_2$</span> correspondence"</a> by Heydeman, Marcolli, Saberi and Stoica,</li> <li><a href="https://arxiv.org/abs/1612.09580" rel="nofollow noreferrer">"Edge length dynamics on graphs with applications to <span class="math-container">$p$</span>-adic AdS/CFT"</a> by Gubser, Heydeman, Jepsen, Marcolli, Parikh, Saberi, Stoica and Trundy. </li> </ol> <p>As another example, in the article <a href="https://link.springer.com/article/10.1007/s00220-019-03665-4" rel="nofollow noreferrer">"A second-quantized Kolmogorov-Chentsov theorem via the operator product expansion"</a> I discovered a general result for making sense of products of random Schwartz distributions starting from Wilson's operator product expansion. The statement of the theorem and the method of proof was done over <span class="math-container">$\mathbb{Q}_p$</span> first.</p> <p>Another remark is that exotic numbers like <span class="math-container">$p$</span>-adics can appear in two different conceptual roles. Suppose one is interested in a "scalar field" <span class="math-container">$\phi:X\rightarrow Y$</span> where <span class="math-container">$X$</span> is space-time. The toy models I mentioned make <span class="math-container">$X$</span> <span class="math-container">$p$</span>-adic but keep <span class="math-container">$Y$</span> real or complex. As in Dyson's quote from Carlo's answer, things should still be Archimedean-valued. This kind of distinction is well known to researchers in the area called the Langlands Program which splits in two separate arenas depending on whether one's favorite <span class="math-container">$L$</span> or zeta functions take values in <span class="math-container">$\mathbb{C}$</span> or <span class="math-container">$\mathbb{Q}_p$</span>. There are works in <span class="math-container">$p$</span>-adic physics (e.g. by Khrennikov) where <span class="math-container">$Y$</span> is taken <span class="math-container">$p$</span>-adic, but I am not familiar with the motivation behind this.</p> <p>Finally, and in relation to Laie's comment above, one possible use of <span class="math-container">$p$</span>-adics is to try to exploit an Adelic product formula in order to compute a real quantity of interest as the inverse of the product of similar quantitites over <span class="math-container">$\mathbb{Q}_p$</span>. This was observed for tree level Veneziano amplitudes and was one of the first motivations for developing <span class="math-container">$p$</span>-adic physics. However, these adelic formulas break down for higher loop amplitudes. See <a href="http://www.sciencedirect.com/science/article/pii/037015739390043D" rel="nofollow noreferrer">this Physics Report article</a> for more on this story.</p> Answer 5: <p>For an overview of applications of p-adic numbers in physics I would refer to the <A HREF="https://en.wikipedia.org/wiki/P-adic_quantum_mechanics">Wikipedia</A> and <A HREF="http://physics.stackexchange.com/questions/15252/p-adic-quantum-mechanic">Physics.stackexchange</A> links, and to this <A HREF="https://ncatlab.org/nlab/show/p-adic+physics">nLab entry.</A> Regarding the second question "<em>What is the most convincing justification in physics that we need to work over the field of real or complex numbers</em>" I would like to quote Freeman Dyson in <A HREF="http://www.ams.org/notices/200902/rtx090200212p.pdf">Birds and Frogs:</A></p> <blockquote> <p>When I look at the history of mathematics, I see a succession of illogical jumps, improbable coincidences, jokes of nature. One of the most profound jokes of nature is the square root of minus one that the physicist Erwin Schrödinger put into his wave equation when he invented wave mechanics in 1926. Starting from wave optics as a model, he wrote down a differential equation for a mechanical particle, but the equation made no sense. The equation looked like the equation of conduction of heat in a continuous medium. Heat conduction has no visible relevance to particle mechanics. Schrödinger’s idea seemed to be going nowhere. But then came the surprise. Schrödinger put the square root of minus one into the equation, and suddenly it made sense. Suddenly it became a wave equation instead of a heat conduction equation. And Schrödinger found to his delight that the equation has solutions corresponding to the quantized orbits in the Bohr model of the atom.</p> <p>It turns out that the Schrödinger equation describes correctly everything we know about the behavior of atoms. It is the basis of all of chemistry and most of physics. <strong>And that square root of minus one means that nature works with complex numbers and not with real numbers.</strong></p> <p>All through the nineteenth century, mathematicians from Abel to Riemann and Weierstrass had been creating a magnificent theory of functions of complex variables. They had discovered that the theory of functions became far deeper and more powerful when it was extended from real to complex numbers. But they always thought of complex numbers as an artificial construction, invented by human mathematicians as a useful and elegant abstraction from real life. It never entered their heads that this artificial number system that they had invented was in fact the ground on which atoms move. They never imagined that nature had got there first.</p> </blockquote> Answer 6: <p>Since the <a href="https://physics.stackexchange.com/questions/15252/p-adic-quantum-mechanic">question on physics.stackexchange</a> that was referred to in the comments above has been closed, let me essentially repeat my answer here together with some updates.</p> <hr> <p>I think there are mainly two reasons for motivating the introduction of <span class="math-container">$p$</span>-adic models in physics.</p> <ol> <li>They could exist in nature.</li> <li>They provide insightful toy models for physical phenomena.</li> </ol> <p>There is a vast physical literature on the subject which cite Reason 1 as justification, this is sometimes called the Vladimirov Hypothesis. Namely, we do not know the texture of spacetime at the Planck scale therefore it is possible that it might look more like <span class="math-container">$\mathbb{Q}_p^d$</span> than <span class="math-container">$\mathbb{R}^d$</span>. It is a seductive idea, but there is no evidence for it. Moreover, this would beg the question of why would Nature choose a particular prime number. A perhaps better hypothesis is the Adelic one advocated by Manin, i.e., all primes should occur democratically. In any case, this is quite speculative at present.</p> <p>The way I see it, <span class="math-container">$p$</span>-adic numbers are useful in physics mainly because of Reason 2. By exaggerating only a little, one could argue that <span class="math-container">$p$</span>-adic toy models is what guided Kenneth Wilson when making his great discoveries in the theory of the renormalization group which entirely revolutionized physics in the early seventies. </p> <p>When studying complex multiscale phenomena it is often important to decompose functions into time-frequency atoms which live on a tree, e.g., when using a wavelet decomposition. Unfortunately for most questions of interest the metric which governs how these atoms interact with each other is not the natural (from the tree point of view) ultrametric distance, but the Euclidean metric of the underlying continuum. Hierarchical models in physics amount to changing the model so it is the ultrametric distance which defines atomic interactions. The same idea also appears in mathematics where such toy models are often called "dyadic models". See <a href="https://terrytao.wordpress.com/2007/07/27/dyadic-models/" rel="nofollow noreferrer">this wonderful post</a> by Tao for a nice discussion of this circle of ideas. Given a problem in Euclidean space, there are lots of ways of setting up a simplified hierarchical model for it. The <span class="math-container">$p$</span>-adics, in some sense provide the most canonical, structured and principled way of doing this.</p> <p>To go back to Wilson and the RG, he also used a hierarchical model and gave it yet another name "the approximate recursion". The importance of this toy model in Wilson's path to discovery is clear from his quote:</p> <blockquote> <p>"Then, at Michael's urging, I work out what happens near four dimensions for the approximate recursion formula, and find that d-4 acts as a small parameter. Knowing this it is then trivial, given my field theoretic training, to construct the beginning of the epsilon expansion for critical exponents."</p> </blockquote> <p>which can be found <a href="http://authors.library.caltech.edu/5456/1/hrst.mit.edu/hrs/renormalization/Wilson/Wilson2.htm" rel="nofollow noreferrer">here</a> (about one third from the bottom of the page).</p> <p>For example, the use of <span class="math-container">$p$</span>-adics gives a toy model for rigorously understanding conformal field theories. This is explained in my article <a href="https://arxiv.org/abs/1511.03180" rel="nofollow noreferrer">"Towards three-dimensional conformal probability"</a> (<a href="http://people.virginia.edu/~aa4cr/TalkUMDfinal.pdf" rel="nofollow noreferrer">these slides</a> are perhaps easier to understand since they have lots of pictures). More recently, the following articles explored similar issues:</p> <ol> <li><a href="https://arxiv.org/abs/1605.01061" rel="nofollow noreferrer">"<span class="math-container">$p$</span>-adic AdS/CFT"</a> by Gubser, Knaute, Parikh, Samberg and Witaszczyk,</li> <li><a href="https://arxiv.org/abs/1605.07639" rel="nofollow noreferrer">"Tensor networks, <span class="math-container">$p$</span>-adic fields, and algebraic curves: arithmetic and the <span class="math-container">${\rm AdS}_3$</span>/<span class="math-container">${\rm CFT}_2$</span> correspondence"</a> by Heydeman, Marcolli, Saberi and Stoica,</li> <li><a href="https://arxiv.org/abs/1612.09580" rel="nofollow noreferrer">"Edge length dynamics on graphs with applications to <span class="math-container">$p$</span>-adic AdS/CFT"</a> by Gubser, Heydeman, Jepsen, Marcolli, Parikh, Saberi, Stoica and Trundy. </li> </ol> <p>As another example, in the article <a href="https://link.springer.com/article/10.1007/s00220-019-03665-4" rel="nofollow noreferrer">"A second-quantized Kolmogorov-Chentsov theorem via the operator product expansion"</a> I discovered a general result for making sense of products of random Schwartz distributions starting from Wilson's operator product expansion. The statement of the theorem and the method of proof was done over <span class="math-container">$\mathbb{Q}_p$</span> first.</p> <p>Another remark is that exotic numbers like <span class="math-container">$p$</span>-adics can appear in two different conceptual roles. Suppose one is interested in a "scalar field" <span class="math-container">$\phi:X\rightarrow Y$</span> where <span class="math-container">$X$</span> is space-time. The toy models I mentioned make <span class="math-container">$X$</span> <span class="math-container">$p$</span>-adic but keep <span class="math-container">$Y$</span> real or complex. As in Dyson's quote from Carlo's answer, things should still be Archimedean-valued. This kind of distinction is well known to researchers in the area called the Langlands Program which splits in two separate arenas depending on whether one's favorite <span class="math-container">$L$</span> or zeta functions take values in <span class="math-container">$\mathbb{C}$</span> or <span class="math-container">$\mathbb{Q}_p$</span>. There are works in <span class="math-container">$p$</span>-adic physics (e.g. by Khrennikov) where <span class="math-container">$Y$</span> is taken <span class="math-container">$p$</span>-adic, but I am not familiar with the motivation behind this.</p> <p>Finally, and in relation to Laie's comment above, one possible use of <span class="math-container">$p$</span>-adics is to try to exploit an Adelic product formula in order to compute a real quantity of interest as the inverse of the product of similar quantitites over <span class="math-container">$\mathbb{Q}_p$</span>. This was observed for tree level Veneziano amplitudes and was one of the first motivations for developing <span class="math-container">$p$</span>-adic physics. However, these adelic formulas break down for higher loop amplitudes. See <a href="http://www.sciencedirect.com/science/article/pii/037015739390043D" rel="nofollow noreferrer">this Physics Report article</a> for more on this story.</p>
A variant of the Monge-Cayley-Salmon theorem?
260,619
17
https://mathoverflow.net/questions/260619/a-variant-of-the-monge-cayley-salmon-theorem
<p>Suppose one has a smooth non-degenerate curve $\gamma: [0,1] \to {\bf R}^n$ into Euclidean space (thus $\gamma'$ never vanishes), with the property that the velocity $\gamma'(t)$ and acceleration $\gamma''(t)$ are always parallel, thus $$ \mathrm{dim}\ \mathrm{span}( \gamma'(t), \gamma''(t) ) \leq 1$$ for all $t \in [0,1]$. To avoid technicalities let us assume that $\gamma$ is a polynomial. Then it is easy to see that $\gamma$ actually traverses a straight line (there is no curvature).</p> <p>More generally, if we have $$ \mathrm{dim}\ \mathrm{span}( \gamma'(t), \dots, \gamma^{(k)}(t) ) \leq k-1$$ for some fixed $k$ and all $t$, then one can show that the curve $\gamma([0,1])$ is trapped inside some (affine) $k-1$-dimensional subspace of ${\bf R}^n$. (For instance, the $k=2$ case corresponds to the case when the curve has no torsion and is thus a plane curve.) One can prove this for instance by showing that the $k$-form $\gamma'(t) \wedge \dots \wedge \gamma^{(k-1)}(t)$ has a direction that is fixed in $t$ (assuming it does not degenerate to zero, in which case one can instead induct).</p> <p>A variant of this is the Monge-Cayley-Salmon theorem: if $\phi: [0,1]^2 \to {\bf R}^3$ is a polynomial immersed surface in three dimensions (thus $D\phi$ has maximal rank everywhere), and for each point $(s,t) \in (0,1)^2$ there is a smooth non-degenerate curve $\gamma = \gamma_{s,t}: [0,1] \to \phi([0,1]^2)$ on the surface passing through $\phi(s,t)$ at time zero, thus $\gamma(0) = \phi(s,t)$ and $\gamma'(0) \neq 0$, which is linear to third order, in that $$ \mathrm{dim}\ \mathrm{span}( \gamma'(0), \gamma''(0), \gamma'''(0) ) \leq 1$$ (or equivalently, $\phi(s,t)$ is a <em>flecnode</em> of the surface, in that there is a tangent line that matches the surface to third order). Then the Monge-Cayley-Salmon theorem asserts that $\phi([0,1]^2)$ is a ruled surface. (I discuss this theorem in <a href="https://terrytao.wordpress.com/2014/03/28/the-cayley-salmon-theorem-via-classical-differential-geometry/">this blog post</a>; the previous results about curves are used in the proof of the Monge-Cayley-Salmon theorem.)</p> <p>In all of the above results, the hypothesis is about the local geometry of the surface (a condition on some finite Taylor expansion of $\phi$, or equivalently some finite jet of the surface), but the conclusion constrains the global geometry of the surface (in particular locating linear spaces that globally relate to the surface).</p> <p>My (somewhat ill-posed) question concerns the following variant of the Monge-Cayley-Salmon situation: suppose that $\phi: [0,1]^2 \to {\bf R}^4$ is a polynomial immersed surface, and suppose that one has the dimension condition $$ \mathrm{dim}\ \mathrm{span}( \phi_s(s,t), \phi_t(s,t), \phi_{ss}(s,t), \phi_{st}(s,t), \phi_{tt}(s,t) ) \leq 3$$ on the first and second partial derivatives of $\phi$ for all $(s,t) \in [0,1]^2$ (thus the Taylor expansion to second order of $\phi([0,1]^2)$ around any point is always at most three-dimensional rather than four). Does this place a strong geometric constraint on the surface $\phi([0,1]^2)$, such as being ruled, or being trapped in a three-dimensional subspace of ${\bf R}^4$? I am a bit vague on what type of conclusion I want here, but it should somehow control the "global" geometry of the surface in a manner similar to the previous examples. The requirement that $\phi$ be a polynomial might be unnecessary, but that is what actually occurs for the application I have in mind.</p> <p>One can also pose this question for higher dimensional varieties in higher dimensional Euclidean spaces, but the above situation of two-dimensional surfaces in ${\bf R}^4$ seems to be the first non-trivial case that is not directly treated by the previous assertions about curves.</p>
[ { "answer_id": 260641, "score": 14, "text": "<p>Setting aside the assumption that $\\phi$ be a polynomial mapping for the moment (however, see below for a construction of a large family of polynomial solutions), if one makes the 'nondegeneracy' assumptions</p>\n\n<ol>\n<li>$\\mathrm{dim}\\ \\mathrm{span}\\bigl( \\phi_s(s,t), \\phi_t(s,t)\\bigr) =2 $,</li>\n<li>$\\mathrm{dim}\\ \\mathrm{span}\\bigl( \\phi_s(s,t), \\phi_t(s,t), \\phi_{ss}(s,t), \\phi_{st}(s,t), \\phi_{tt}(s,t) \\bigr) = 3$ for all $(s,t)\\in[0,1]^2$, and </li>\n<li>the subspace $W(s,t) =\\mathrm{span}\\bigl( \\phi_s(s,t), \\phi_t(s,t), \\phi_{ss}(s,t), \\phi_{st}(s,t), \\phi_{tt}(s,t) \\bigr)\\subset\\mathbb{R}^4$ is not constant, in the sense that $W:[0,1]^2\\to \\mathrm{Gr}_3(\\mathbb{R}^4)\\simeq\\mathbb{RP}^3$ has nonvanishing differential,</li>\n</ol>\n\n<p>then one can show that the surface $\\phi\\bigl([0,1]^2\\bigr)\\subset\\mathbb{R}^4$ is ruled (and does not lie in an affine $3$-space). </p>\n\n<p>Such surfaces locally depend on three arbitrary functions of one variable in Cartan's sense. One way of describing them is this: Let $\\Lambda$ be the space of (affine) lines in $\\mathbb{R}^4$, a manifold of dimension $6$. Consider the ($9$-dimensional) bundle $\\pi:F\\to\\Lambda$ whose fiber over $\\lambda\\in\\Lambda$ is the flag variety of the $3$-dimensional vector space $\\mathbb{R}^4/\\lambda'$, where $\\lambda'\\subset\\mathbb{R}^4$ is the linear subspace parallel to $\\lambda$. Then there exists a smooth $4$-plane field $D\\subset TF$ such that, if $\\gamma\\subset F$ is a generic curve tangent to $D$, then regarding $\\gamma$ as a $1$-parameter family of affine lines via $\\pi(\\gamma)\\subset \\Lambda$, the union of these lines is a surface satisfying the above assumptions. (Here 'generic' means that the tangents to $\\gamma$ do not lie in triple of hyperplanes in $D$.)</p>\n\n<p>It is not hard to write down polynomial solutions: For example, if $f:\\mathbb{R}\\to\\mathbb{R}^4$ is a polynomial curve that satisfies $f'(s)\\wedge f''(s)\\wedge f'''(s) \\wedge f''''(s)\\not = 0$, then the mapping $\\phi:\\mathbb{R}^2\\to\\mathbb{R}^4$ given by \n$$\n\\phi(s,t) = f(s) + f'(s)\\, t\n$$\n(which parametrizes the 'tangential development' of the curve $f$) satisfies these conditions when $t\\not=0$. (By replacing $t$ by $t{+}1$, say, one could arrange that $\\phi$ be an immersion on all of $[0,1]^2$.)</p>\n\n<p>The following analysis is a more-or-less standard approach to verifying the above description using the so-called 'moving frame'. (I'm sure that the result itself\nis classical in some sense, though I don't know offhand where to look\nin the literature to find it.)</p>\n\n<p>The three conditions listed above are actually independent of the choice of $st$-coordinates on the surface in $\\mathbb{R}^4$ and so can be regarded as conditions on a surface $S\\subset\\mathbb{R}^4$ that, for local analysis purposes, can be taken to be smoothly embedded.</p>\n\n<p>Let $B_0(S)$ denote the space of quintuples $(p;v_1,v_2,v_3,v_4)$ where $p$ lies in $S$ and the quadruple $(v_1,v_2,v_3,v_4)$ is a basis of $\\mathbb{R}^4$ such that $(v_1,v_2)$ is a basis of $T_pS$ while $(v_1,v_2,v_3)$ is a basis of the $3$-dimensional subspace $W_pS\\subset \\mathbb{R}^4$. Then $B_0$ is a smooth submanifold of the product $\\mathbb{R}^4\\times\\mathrm{GL}(4,\\mathbb{R})$ that has dimension $2 + 4 + 3 + 4 = 13$. \n<p> \nIt is useful to define $\\mathbb{R}^4$-valued functions $x,e_i: \\mathbb{R}^4\\times\\mathrm{GL}(4,\\mathbb{R})\\to\\mathbb{R}^4$ such that $x(p;v_1,v_2,v_3,v_4) = p$ while $e_i(p;v_1,v_2,v_3,v_4) = v_i$. Then there\nexist unique (linearly independent) $1$-forms $\\omega^i$ and $\\theta^i_j$ \non $\\mathbb{R}^4\\times\\mathrm{GL}(4,\\mathbb{R})$ such that (assuming the\nusual summation convention on repeated indices) the following <em>structure equations</em> hold:\n$$\n\\mathrm{d}x = e_i \\omega^i\n\\qquad\\text{and}\\qquad\n\\mathrm{d}e_i = e_j\\,\\theta^j_i\\,.\\tag1\n$$ </p>\n\n<p>$$\n\\mathrm{d}\\omega^i = -\\theta^i_j\\wedge\\omega^j\n\\qquad\\text{and}\\qquad\n\\mathrm{d}\\theta^i_j = -\\theta^i_k\\wedge\\theta^k_j\\,.\\tag2\n$$</p>\n\n<p>Now, pull back these functions and $1$-forms to $B_0(S)$ (but, as is customary, not notate the pullback). The definition of $B_0(S)$ and the assumptions on $S$ imply that\n$$\n\\mathrm{d}x\\wedge e_1\\wedge e_2 = \\mathrm{d}e_1\\wedge e_1\\wedge e_2\\wedge e_3\n= \\mathrm{d}e_2\\wedge e_1\\wedge e_2\\wedge e_3 = 0\n$$\nwhile the two expressions \n$$\n\\bigl(\\,\\mathrm{d}e_1\\wedge e_1\\wedge e_2,\\ \\mathrm{d}e_2\\wedge e_1\\wedge e_2\\bigr)\n\\qquad\\text{and}\\qquad\n\\mathrm{d}e_3\\wedge e_1\\wedge e_2\\wedge e_3\n$$\nare nowhere vanishing. Using the above structure equations, these imply the relations\n$$\n\\omega^3 = \\omega^4 = \\theta^4_1 = \\theta^4_2 = 0,\\tag3\n$$\nwhile, because of the three assumptions, $\\omega^1\\wedge\\omega^2$ is nonvanishing, the pair $(\\theta^3_1,\\theta^3_2)$ do not simultaneously vanish,\nand $\\theta^4_3$ is nonvanishing.</p>\n\n<p>Meanwhile, the structure equations yield\n$$\n0 = \\mathrm{d}\\omega^3 \n= -\\theta^3_1\\wedge\\omega^1 -\\theta^3_2\\wedge\\omega^2,\n$$\nso there must exist functions $h_{ij}=h_{ji}$ for $1\\le i,j,\\le 2$, not all simultaneously vanishing, such that $\\theta^3_i = h_{ij}\\omega^j$. The quadratic form $h = h_{ij}\\,\\omega^i\\omega^j$ is then nonvanishing and well-defined up to multiples on the surface $S$. Moreover, for $i = 1$ or $2$, \n$$\n0 = \\mathrm{d}\\theta^4_i \n= -\\theta^4_k\\wedge\\theta^k_i = -\\theta^4_3\\wedge\\theta^3_i\\,.\n$$\nThus, since $\\theta^4_3$ is nonvanishing, it follows that $\\theta^3_1$ and $\\theta^3_2$ are multiples of $\\theta^4_3$. In particular, $\\theta^3_1\\wedge\\theta^3_2$ vanishes identically, so $h_{11}h_{22}-{h_{12}}^2$ vanishes identically. Thus, the quadratic form $h$ has rank $1$.</p>\n\n<p>Let $B_1(S)\\subset B_2(S)$ denote the submanifold defined by $h_{11} = h_{12}=0$. It is a smooth submanifold of $B_0(S)$ of dimension $11$, and when all the forms and functions are pulled back to $B_1(S)$, we have $\\theta^3_1 = 0$ while $\\theta^3_2 = h_{22}\\,\\omega^2$. In particular, it now follows that $\\theta^4_3$ is also a multiple of $\\omega^2$, say $\\theta^4_3 = f\\,\\omega^2$ for some $f$ (which is nonvanishing).</p>\n\n<p>Moreover,\n$$\n0 = \\mathrm{d}\\theta^3_1 = -\\theta^3_k\\wedge\\theta^k_1 \n= -\\theta^3_2\\wedge \\theta^2_1 = -h_{22}\\,\\omega^2\\wedge\\theta^2_1, \n$$ \nso it follows (since $h_{22}$ is nonvanishing) that $\\theta^2_1 = g\\,\\omega^2$ for some function $g$ on $B_1(S)$. </p>\n\n<p>Now,\n$$\n\\mathrm{d}\\omega^2 = -\\theta^2_j\\wedge\\omega^j \n= -g\\,\\omega^2\\wedge\\omega^1 - \\theta^2_2\\wedge\\omega^2\n= -(\\theta^2_2 - g\\,\\omega^1)\\wedge\\omega^2.\n$$ \nThus, $\\omega^2$ is an integrable $1$-form, and, because it is semi-basic\nfor the submersion $x:B_1(S)\\to S\\subset\\mathbb{R}^4$, it follows that $\\omega^2$ is a multiple of the $x$-pullback of a (nonvanishing) $1$-form on $S$. Thus, $S$ is foliated by (connected) curves whose $x$-preimages in $B_1(S)$ are codimension $1$ integral submanifolds of $\\omega^2$.</p>\n\n<p>I claim that these curves in $S$ are, in fact, lines in $\\mathbb{R}^4$. To see this, note that, when one pulls back to a leaf of $\\omega^2=0$ in $B_1(S)$, one has $\\theta^2_1 = \\theta^3_1=\\theta^4_1= 0$ as well, so one has\n$$\n\\mathrm{d}x = e_1\\,\\omega^1\\qquad\\text{and}\\qquad \\mathrm{d}e_1 = e_1\\,\\theta^1_1\\,.\n$$\nIn particular, the direction of $e_1$ is fixed on this leaf, and this is the tangent direction of the mapping $x$ restricted to this leaf. Hence the $x$-image of this leaf is an open interval in a line in $\\mathbb{R}^4$. Thus, the surface is ruled, as claimed.<p></p>\n\n<p>There still remains the question of how one could 'generate' these surfaces, at least locally. I claim that, since these surfaces are, in fact, $1$-parameter families of lines, one should think of them as curves in the space of lines that satisfy some differential equations. Here is how one can think of this system of equations:<p></p>\n\n<p>Let $\\mathcal{I}$ denote the Pfaffian system of rank $5$ that is generated by the five linearly independent $1$-forms\n$$\n\\omega^3,\\ \\omega^4,\\ \\theta^4_1\\,,\\ \\theta^4_2\\,,\\ \\theta^3_1 \n$$\n(these are the $1$-forms that vanish when pulled back to $B_1(S)$ when $S\\subset\\mathbb{R}^4$ is a surface satisfying our hypotheses). By the structure equations, \n$$\n\\left.\n\\begin{aligned}\n\\mathrm{d}\\omega^4 &amp;\\equiv 0\\\\\n\\mathrm{d}\\theta^4_1 &amp;\\equiv 0\\\\\n\\mathrm{d}\\omega^3 &amp;\\equiv -\\theta^3_2\\wedge\\omega^2\\\\\n\\mathrm{d}\\theta^3_1 &amp;\\equiv -\\theta^3_2\\wedge\\theta^2_1\\\\\n\\mathrm{d}\\theta^4_2 &amp;\\equiv \\phantom{-}\\theta^3_2\\wedge\\theta^4_3\\\\\n\\end{aligned}\\ \\right\\} \\mathrm{modulo}\\ \\mathcal{I}\n$$\nIt follows that $\\mathcal{J}$, the Cartan system of $\\mathcal{I}$, \nhas rank $9$ and is spanned by the nine $1$-forms\n$$\n\\omega^3,\\ \\omega^4,\\ \\theta^4_1\\,,\\ \\theta^4_2\\,,\\ \\theta^3_1\\,,\\ \n\\omega^2,\\ \\theta^2_1\\,,\\ \\theta^3_2\\,,\\ \\theta^4_3\n$$\nIn fact, $\\mathcal{J}$ is easily seen to be the bundle of $1$-forms \non $\\mathbb{R}^4\\times\\mathrm{GL}(4,\\mathbb{R})$ that is semibasic for the projection to the $9$-manifold $F$, where the projection sends $(p; v_i)$\nto the element of $F$ described by\n$$\n\\bigl (\\lambda(p,v_1),\\ [v_1\\wedge v_2], [v_1\\wedge v_2\\wedge v_3]\\ \\bigr)\n$$\nand where $\\lambda(p,v_1)$ is the line through $p$ in the direction $v_1$.</p>\n\n<p>In particular $\\mathcal{I}$ is the pullback of a well-defined Pfaffian system of rank $5$ on $F$, whose annihilator is the $4$-plane field $D\\subset TF$. Moreover, it is easy to show that the three rank $6$ Pfaffian systems generated by adjoining any one of $\\omega^2$, $\\theta^3_2$, or $\\theta^4_3$ to $\\mathcal{I}$ are themselves pullbacks of rank $6$ Pfaffian systems on $F$ whose annihilators in $TF$ are each $3$-plane subbundles of $D$. <p>\nBy its very construction, the projection of $B_1(S)$ into $F$ is a curve that is tangent to $D$ and not tangent to any of these three $3$-plane subbundles of $D$.</p>\n\n<p>Conversely, if $\\gamma\\subset F$ is any integral curve of $D$ that is not tangent to any of the three $3$-plane subbundles of $D$, its preimage in $\\mathbb{R}^4\\times\\mathrm{GL}(4,\\mathbb{R})$ is a submanifold of the form $B_1(S)$ for a surface $S$ satisfying our conditions, in fact, the surface swept out by the union of the lines represented by $\\lambda(\\gamma)$, where $\\lambda:F\\to \\Lambda$ is the obvious map to the lines.</p>\n\n<p>It is a standard fact that the curves tangent to a 4-plane field in a smooth manifold are locally described (up to reparametrization) by prescribing $3$ functions of one variable.</p>\n" }, { "answer_id": 260641, "score": 14, "text": "<p>Setting aside the assumption that $\\phi$ be a polynomial mapping for the moment (however, see below for a construction of a large family of polynomial solutions), if one makes the 'nondegeneracy' assumptions</p>\n\n<ol>\n<li>$\\mathrm{dim}\\ \\mathrm{span}\\bigl( \\phi_s(s,t), \\phi_t(s,t)\\bigr) =2 $,</li>\n<li>$\\mathrm{dim}\\ \\mathrm{span}\\bigl( \\phi_s(s,t), \\phi_t(s,t), \\phi_{ss}(s,t), \\phi_{st}(s,t), \\phi_{tt}(s,t) \\bigr) = 3$ for all $(s,t)\\in[0,1]^2$, and </li>\n<li>the subspace $W(s,t) =\\mathrm{span}\\bigl( \\phi_s(s,t), \\phi_t(s,t), \\phi_{ss}(s,t), \\phi_{st}(s,t), \\phi_{tt}(s,t) \\bigr)\\subset\\mathbb{R}^4$ is not constant, in the sense that $W:[0,1]^2\\to \\mathrm{Gr}_3(\\mathbb{R}^4)\\simeq\\mathbb{RP}^3$ has nonvanishing differential,</li>\n</ol>\n\n<p>then one can show that the surface $\\phi\\bigl([0,1]^2\\bigr)\\subset\\mathbb{R}^4$ is ruled (and does not lie in an affine $3$-space). </p>\n\n<p>Such surfaces locally depend on three arbitrary functions of one variable in Cartan's sense. One way of describing them is this: Let $\\Lambda$ be the space of (affine) lines in $\\mathbb{R}^4$, a manifold of dimension $6$. Consider the ($9$-dimensional) bundle $\\pi:F\\to\\Lambda$ whose fiber over $\\lambda\\in\\Lambda$ is the flag variety of the $3$-dimensional vector space $\\mathbb{R}^4/\\lambda'$, where $\\lambda'\\subset\\mathbb{R}^4$ is the linear subspace parallel to $\\lambda$. Then there exists a smooth $4$-plane field $D\\subset TF$ such that, if $\\gamma\\subset F$ is a generic curve tangent to $D$, then regarding $\\gamma$ as a $1$-parameter family of affine lines via $\\pi(\\gamma)\\subset \\Lambda$, the union of these lines is a surface satisfying the above assumptions. (Here 'generic' means that the tangents to $\\gamma$ do not lie in triple of hyperplanes in $D$.)</p>\n\n<p>It is not hard to write down polynomial solutions: For example, if $f:\\mathbb{R}\\to\\mathbb{R}^4$ is a polynomial curve that satisfies $f'(s)\\wedge f''(s)\\wedge f'''(s) \\wedge f''''(s)\\not = 0$, then the mapping $\\phi:\\mathbb{R}^2\\to\\mathbb{R}^4$ given by \n$$\n\\phi(s,t) = f(s) + f'(s)\\, t\n$$\n(which parametrizes the 'tangential development' of the curve $f$) satisfies these conditions when $t\\not=0$. (By replacing $t$ by $t{+}1$, say, one could arrange that $\\phi$ be an immersion on all of $[0,1]^2$.)</p>\n\n<p>The following analysis is a more-or-less standard approach to verifying the above description using the so-called 'moving frame'. (I'm sure that the result itself\nis classical in some sense, though I don't know offhand where to look\nin the literature to find it.)</p>\n\n<p>The three conditions listed above are actually independent of the choice of $st$-coordinates on the surface in $\\mathbb{R}^4$ and so can be regarded as conditions on a surface $S\\subset\\mathbb{R}^4$ that, for local analysis purposes, can be taken to be smoothly embedded.</p>\n\n<p>Let $B_0(S)$ denote the space of quintuples $(p;v_1,v_2,v_3,v_4)$ where $p$ lies in $S$ and the quadruple $(v_1,v_2,v_3,v_4)$ is a basis of $\\mathbb{R}^4$ such that $(v_1,v_2)$ is a basis of $T_pS$ while $(v_1,v_2,v_3)$ is a basis of the $3$-dimensional subspace $W_pS\\subset \\mathbb{R}^4$. Then $B_0$ is a smooth submanifold of the product $\\mathbb{R}^4\\times\\mathrm{GL}(4,\\mathbb{R})$ that has dimension $2 + 4 + 3 + 4 = 13$. \n<p> \nIt is useful to define $\\mathbb{R}^4$-valued functions $x,e_i: \\mathbb{R}^4\\times\\mathrm{GL}(4,\\mathbb{R})\\to\\mathbb{R}^4$ such that $x(p;v_1,v_2,v_3,v_4) = p$ while $e_i(p;v_1,v_2,v_3,v_4) = v_i$. Then there\nexist unique (linearly independent) $1$-forms $\\omega^i$ and $\\theta^i_j$ \non $\\mathbb{R}^4\\times\\mathrm{GL}(4,\\mathbb{R})$ such that (assuming the\nusual summation convention on repeated indices) the following <em>structure equations</em> hold:\n$$\n\\mathrm{d}x = e_i \\omega^i\n\\qquad\\text{and}\\qquad\n\\mathrm{d}e_i = e_j\\,\\theta^j_i\\,.\\tag1\n$$ </p>\n\n<p>$$\n\\mathrm{d}\\omega^i = -\\theta^i_j\\wedge\\omega^j\n\\qquad\\text{and}\\qquad\n\\mathrm{d}\\theta^i_j = -\\theta^i_k\\wedge\\theta^k_j\\,.\\tag2\n$$</p>\n\n<p>Now, pull back these functions and $1$-forms to $B_0(S)$ (but, as is customary, not notate the pullback). The definition of $B_0(S)$ and the assumptions on $S$ imply that\n$$\n\\mathrm{d}x\\wedge e_1\\wedge e_2 = \\mathrm{d}e_1\\wedge e_1\\wedge e_2\\wedge e_3\n= \\mathrm{d}e_2\\wedge e_1\\wedge e_2\\wedge e_3 = 0\n$$\nwhile the two expressions \n$$\n\\bigl(\\,\\mathrm{d}e_1\\wedge e_1\\wedge e_2,\\ \\mathrm{d}e_2\\wedge e_1\\wedge e_2\\bigr)\n\\qquad\\text{and}\\qquad\n\\mathrm{d}e_3\\wedge e_1\\wedge e_2\\wedge e_3\n$$\nare nowhere vanishing. Using the above structure equations, these imply the relations\n$$\n\\omega^3 = \\omega^4 = \\theta^4_1 = \\theta^4_2 = 0,\\tag3\n$$\nwhile, because of the three assumptions, $\\omega^1\\wedge\\omega^2$ is nonvanishing, the pair $(\\theta^3_1,\\theta^3_2)$ do not simultaneously vanish,\nand $\\theta^4_3$ is nonvanishing.</p>\n\n<p>Meanwhile, the structure equations yield\n$$\n0 = \\mathrm{d}\\omega^3 \n= -\\theta^3_1\\wedge\\omega^1 -\\theta^3_2\\wedge\\omega^2,\n$$\nso there must exist functions $h_{ij}=h_{ji}$ for $1\\le i,j,\\le 2$, not all simultaneously vanishing, such that $\\theta^3_i = h_{ij}\\omega^j$. The quadratic form $h = h_{ij}\\,\\omega^i\\omega^j$ is then nonvanishing and well-defined up to multiples on the surface $S$. Moreover, for $i = 1$ or $2$, \n$$\n0 = \\mathrm{d}\\theta^4_i \n= -\\theta^4_k\\wedge\\theta^k_i = -\\theta^4_3\\wedge\\theta^3_i\\,.\n$$\nThus, since $\\theta^4_3$ is nonvanishing, it follows that $\\theta^3_1$ and $\\theta^3_2$ are multiples of $\\theta^4_3$. In particular, $\\theta^3_1\\wedge\\theta^3_2$ vanishes identically, so $h_{11}h_{22}-{h_{12}}^2$ vanishes identically. Thus, the quadratic form $h$ has rank $1$.</p>\n\n<p>Let $B_1(S)\\subset B_2(S)$ denote the submanifold defined by $h_{11} = h_{12}=0$. It is a smooth submanifold of $B_0(S)$ of dimension $11$, and when all the forms and functions are pulled back to $B_1(S)$, we have $\\theta^3_1 = 0$ while $\\theta^3_2 = h_{22}\\,\\omega^2$. In particular, it now follows that $\\theta^4_3$ is also a multiple of $\\omega^2$, say $\\theta^4_3 = f\\,\\omega^2$ for some $f$ (which is nonvanishing).</p>\n\n<p>Moreover,\n$$\n0 = \\mathrm{d}\\theta^3_1 = -\\theta^3_k\\wedge\\theta^k_1 \n= -\\theta^3_2\\wedge \\theta^2_1 = -h_{22}\\,\\omega^2\\wedge\\theta^2_1, \n$$ \nso it follows (since $h_{22}$ is nonvanishing) that $\\theta^2_1 = g\\,\\omega^2$ for some function $g$ on $B_1(S)$. </p>\n\n<p>Now,\n$$\n\\mathrm{d}\\omega^2 = -\\theta^2_j\\wedge\\omega^j \n= -g\\,\\omega^2\\wedge\\omega^1 - \\theta^2_2\\wedge\\omega^2\n= -(\\theta^2_2 - g\\,\\omega^1)\\wedge\\omega^2.\n$$ \nThus, $\\omega^2$ is an integrable $1$-form, and, because it is semi-basic\nfor the submersion $x:B_1(S)\\to S\\subset\\mathbb{R}^4$, it follows that $\\omega^2$ is a multiple of the $x$-pullback of a (nonvanishing) $1$-form on $S$. Thus, $S$ is foliated by (connected) curves whose $x$-preimages in $B_1(S)$ are codimension $1$ integral submanifolds of $\\omega^2$.</p>\n\n<p>I claim that these curves in $S$ are, in fact, lines in $\\mathbb{R}^4$. To see this, note that, when one pulls back to a leaf of $\\omega^2=0$ in $B_1(S)$, one has $\\theta^2_1 = \\theta^3_1=\\theta^4_1= 0$ as well, so one has\n$$\n\\mathrm{d}x = e_1\\,\\omega^1\\qquad\\text{and}\\qquad \\mathrm{d}e_1 = e_1\\,\\theta^1_1\\,.\n$$\nIn particular, the direction of $e_1$ is fixed on this leaf, and this is the tangent direction of the mapping $x$ restricted to this leaf. Hence the $x$-image of this leaf is an open interval in a line in $\\mathbb{R}^4$. Thus, the surface is ruled, as claimed.<p></p>\n\n<p>There still remains the question of how one could 'generate' these surfaces, at least locally. I claim that, since these surfaces are, in fact, $1$-parameter families of lines, one should think of them as curves in the space of lines that satisfy some differential equations. Here is how one can think of this system of equations:<p></p>\n\n<p>Let $\\mathcal{I}$ denote the Pfaffian system of rank $5$ that is generated by the five linearly independent $1$-forms\n$$\n\\omega^3,\\ \\omega^4,\\ \\theta^4_1\\,,\\ \\theta^4_2\\,,\\ \\theta^3_1 \n$$\n(these are the $1$-forms that vanish when pulled back to $B_1(S)$ when $S\\subset\\mathbb{R}^4$ is a surface satisfying our hypotheses). By the structure equations, \n$$\n\\left.\n\\begin{aligned}\n\\mathrm{d}\\omega^4 &amp;\\equiv 0\\\\\n\\mathrm{d}\\theta^4_1 &amp;\\equiv 0\\\\\n\\mathrm{d}\\omega^3 &amp;\\equiv -\\theta^3_2\\wedge\\omega^2\\\\\n\\mathrm{d}\\theta^3_1 &amp;\\equiv -\\theta^3_2\\wedge\\theta^2_1\\\\\n\\mathrm{d}\\theta^4_2 &amp;\\equiv \\phantom{-}\\theta^3_2\\wedge\\theta^4_3\\\\\n\\end{aligned}\\ \\right\\} \\mathrm{modulo}\\ \\mathcal{I}\n$$\nIt follows that $\\mathcal{J}$, the Cartan system of $\\mathcal{I}$, \nhas rank $9$ and is spanned by the nine $1$-forms\n$$\n\\omega^3,\\ \\omega^4,\\ \\theta^4_1\\,,\\ \\theta^4_2\\,,\\ \\theta^3_1\\,,\\ \n\\omega^2,\\ \\theta^2_1\\,,\\ \\theta^3_2\\,,\\ \\theta^4_3\n$$\nIn fact, $\\mathcal{J}$ is easily seen to be the bundle of $1$-forms \non $\\mathbb{R}^4\\times\\mathrm{GL}(4,\\mathbb{R})$ that is semibasic for the projection to the $9$-manifold $F$, where the projection sends $(p; v_i)$\nto the element of $F$ described by\n$$\n\\bigl (\\lambda(p,v_1),\\ [v_1\\wedge v_2], [v_1\\wedge v_2\\wedge v_3]\\ \\bigr)\n$$\nand where $\\lambda(p,v_1)$ is the line through $p$ in the direction $v_1$.</p>\n\n<p>In particular $\\mathcal{I}$ is the pullback of a well-defined Pfaffian system of rank $5$ on $F$, whose annihilator is the $4$-plane field $D\\subset TF$. Moreover, it is easy to show that the three rank $6$ Pfaffian systems generated by adjoining any one of $\\omega^2$, $\\theta^3_2$, or $\\theta^4_3$ to $\\mathcal{I}$ are themselves pullbacks of rank $6$ Pfaffian systems on $F$ whose annihilators in $TF$ are each $3$-plane subbundles of $D$. <p>\nBy its very construction, the projection of $B_1(S)$ into $F$ is a curve that is tangent to $D$ and not tangent to any of these three $3$-plane subbundles of $D$.</p>\n\n<p>Conversely, if $\\gamma\\subset F$ is any integral curve of $D$ that is not tangent to any of the three $3$-plane subbundles of $D$, its preimage in $\\mathbb{R}^4\\times\\mathrm{GL}(4,\\mathbb{R})$ is a submanifold of the form $B_1(S)$ for a surface $S$ satisfying our conditions, in fact, the surface swept out by the union of the lines represented by $\\lambda(\\gamma)$, where $\\lambda:F\\to \\Lambda$ is the obvious map to the lines.</p>\n\n<p>It is a standard fact that the curves tangent to a 4-plane field in a smooth manifold are locally described (up to reparametrization) by prescribing $3$ functions of one variable.</p>\n" }, { "answer_id": 260641, "score": 14, "text": "<p>Setting aside the assumption that $\\phi$ be a polynomial mapping for the moment (however, see below for a construction of a large family of polynomial solutions), if one makes the 'nondegeneracy' assumptions</p>\n\n<ol>\n<li>$\\mathrm{dim}\\ \\mathrm{span}\\bigl( \\phi_s(s,t), \\phi_t(s,t)\\bigr) =2 $,</li>\n<li>$\\mathrm{dim}\\ \\mathrm{span}\\bigl( \\phi_s(s,t), \\phi_t(s,t), \\phi_{ss}(s,t), \\phi_{st}(s,t), \\phi_{tt}(s,t) \\bigr) = 3$ for all $(s,t)\\in[0,1]^2$, and </li>\n<li>the subspace $W(s,t) =\\mathrm{span}\\bigl( \\phi_s(s,t), \\phi_t(s,t), \\phi_{ss}(s,t), \\phi_{st}(s,t), \\phi_{tt}(s,t) \\bigr)\\subset\\mathbb{R}^4$ is not constant, in the sense that $W:[0,1]^2\\to \\mathrm{Gr}_3(\\mathbb{R}^4)\\simeq\\mathbb{RP}^3$ has nonvanishing differential,</li>\n</ol>\n\n<p>then one can show that the surface $\\phi\\bigl([0,1]^2\\bigr)\\subset\\mathbb{R}^4$ is ruled (and does not lie in an affine $3$-space). </p>\n\n<p>Such surfaces locally depend on three arbitrary functions of one variable in Cartan's sense. One way of describing them is this: Let $\\Lambda$ be the space of (affine) lines in $\\mathbb{R}^4$, a manifold of dimension $6$. Consider the ($9$-dimensional) bundle $\\pi:F\\to\\Lambda$ whose fiber over $\\lambda\\in\\Lambda$ is the flag variety of the $3$-dimensional vector space $\\mathbb{R}^4/\\lambda'$, where $\\lambda'\\subset\\mathbb{R}^4$ is the linear subspace parallel to $\\lambda$. Then there exists a smooth $4$-plane field $D\\subset TF$ such that, if $\\gamma\\subset F$ is a generic curve tangent to $D$, then regarding $\\gamma$ as a $1$-parameter family of affine lines via $\\pi(\\gamma)\\subset \\Lambda$, the union of these lines is a surface satisfying the above assumptions. (Here 'generic' means that the tangents to $\\gamma$ do not lie in triple of hyperplanes in $D$.)</p>\n\n<p>It is not hard to write down polynomial solutions: For example, if $f:\\mathbb{R}\\to\\mathbb{R}^4$ is a polynomial curve that satisfies $f'(s)\\wedge f''(s)\\wedge f'''(s) \\wedge f''''(s)\\not = 0$, then the mapping $\\phi:\\mathbb{R}^2\\to\\mathbb{R}^4$ given by \n$$\n\\phi(s,t) = f(s) + f'(s)\\, t\n$$\n(which parametrizes the 'tangential development' of the curve $f$) satisfies these conditions when $t\\not=0$. (By replacing $t$ by $t{+}1$, say, one could arrange that $\\phi$ be an immersion on all of $[0,1]^2$.)</p>\n\n<p>The following analysis is a more-or-less standard approach to verifying the above description using the so-called 'moving frame'. (I'm sure that the result itself\nis classical in some sense, though I don't know offhand where to look\nin the literature to find it.)</p>\n\n<p>The three conditions listed above are actually independent of the choice of $st$-coordinates on the surface in $\\mathbb{R}^4$ and so can be regarded as conditions on a surface $S\\subset\\mathbb{R}^4$ that, for local analysis purposes, can be taken to be smoothly embedded.</p>\n\n<p>Let $B_0(S)$ denote the space of quintuples $(p;v_1,v_2,v_3,v_4)$ where $p$ lies in $S$ and the quadruple $(v_1,v_2,v_3,v_4)$ is a basis of $\\mathbb{R}^4$ such that $(v_1,v_2)$ is a basis of $T_pS$ while $(v_1,v_2,v_3)$ is a basis of the $3$-dimensional subspace $W_pS\\subset \\mathbb{R}^4$. Then $B_0$ is a smooth submanifold of the product $\\mathbb{R}^4\\times\\mathrm{GL}(4,\\mathbb{R})$ that has dimension $2 + 4 + 3 + 4 = 13$. \n<p> \nIt is useful to define $\\mathbb{R}^4$-valued functions $x,e_i: \\mathbb{R}^4\\times\\mathrm{GL}(4,\\mathbb{R})\\to\\mathbb{R}^4$ such that $x(p;v_1,v_2,v_3,v_4) = p$ while $e_i(p;v_1,v_2,v_3,v_4) = v_i$. Then there\nexist unique (linearly independent) $1$-forms $\\omega^i$ and $\\theta^i_j$ \non $\\mathbb{R}^4\\times\\mathrm{GL}(4,\\mathbb{R})$ such that (assuming the\nusual summation convention on repeated indices) the following <em>structure equations</em> hold:\n$$\n\\mathrm{d}x = e_i \\omega^i\n\\qquad\\text{and}\\qquad\n\\mathrm{d}e_i = e_j\\,\\theta^j_i\\,.\\tag1\n$$ </p>\n\n<p>$$\n\\mathrm{d}\\omega^i = -\\theta^i_j\\wedge\\omega^j\n\\qquad\\text{and}\\qquad\n\\mathrm{d}\\theta^i_j = -\\theta^i_k\\wedge\\theta^k_j\\,.\\tag2\n$$</p>\n\n<p>Now, pull back these functions and $1$-forms to $B_0(S)$ (but, as is customary, not notate the pullback). The definition of $B_0(S)$ and the assumptions on $S$ imply that\n$$\n\\mathrm{d}x\\wedge e_1\\wedge e_2 = \\mathrm{d}e_1\\wedge e_1\\wedge e_2\\wedge e_3\n= \\mathrm{d}e_2\\wedge e_1\\wedge e_2\\wedge e_3 = 0\n$$\nwhile the two expressions \n$$\n\\bigl(\\,\\mathrm{d}e_1\\wedge e_1\\wedge e_2,\\ \\mathrm{d}e_2\\wedge e_1\\wedge e_2\\bigr)\n\\qquad\\text{and}\\qquad\n\\mathrm{d}e_3\\wedge e_1\\wedge e_2\\wedge e_3\n$$\nare nowhere vanishing. Using the above structure equations, these imply the relations\n$$\n\\omega^3 = \\omega^4 = \\theta^4_1 = \\theta^4_2 = 0,\\tag3\n$$\nwhile, because of the three assumptions, $\\omega^1\\wedge\\omega^2$ is nonvanishing, the pair $(\\theta^3_1,\\theta^3_2)$ do not simultaneously vanish,\nand $\\theta^4_3$ is nonvanishing.</p>\n\n<p>Meanwhile, the structure equations yield\n$$\n0 = \\mathrm{d}\\omega^3 \n= -\\theta^3_1\\wedge\\omega^1 -\\theta^3_2\\wedge\\omega^2,\n$$\nso there must exist functions $h_{ij}=h_{ji}$ for $1\\le i,j,\\le 2$, not all simultaneously vanishing, such that $\\theta^3_i = h_{ij}\\omega^j$. The quadratic form $h = h_{ij}\\,\\omega^i\\omega^j$ is then nonvanishing and well-defined up to multiples on the surface $S$. Moreover, for $i = 1$ or $2$, \n$$\n0 = \\mathrm{d}\\theta^4_i \n= -\\theta^4_k\\wedge\\theta^k_i = -\\theta^4_3\\wedge\\theta^3_i\\,.\n$$\nThus, since $\\theta^4_3$ is nonvanishing, it follows that $\\theta^3_1$ and $\\theta^3_2$ are multiples of $\\theta^4_3$. In particular, $\\theta^3_1\\wedge\\theta^3_2$ vanishes identically, so $h_{11}h_{22}-{h_{12}}^2$ vanishes identically. Thus, the quadratic form $h$ has rank $1$.</p>\n\n<p>Let $B_1(S)\\subset B_2(S)$ denote the submanifold defined by $h_{11} = h_{12}=0$. It is a smooth submanifold of $B_0(S)$ of dimension $11$, and when all the forms and functions are pulled back to $B_1(S)$, we have $\\theta^3_1 = 0$ while $\\theta^3_2 = h_{22}\\,\\omega^2$. In particular, it now follows that $\\theta^4_3$ is also a multiple of $\\omega^2$, say $\\theta^4_3 = f\\,\\omega^2$ for some $f$ (which is nonvanishing).</p>\n\n<p>Moreover,\n$$\n0 = \\mathrm{d}\\theta^3_1 = -\\theta^3_k\\wedge\\theta^k_1 \n= -\\theta^3_2\\wedge \\theta^2_1 = -h_{22}\\,\\omega^2\\wedge\\theta^2_1, \n$$ \nso it follows (since $h_{22}$ is nonvanishing) that $\\theta^2_1 = g\\,\\omega^2$ for some function $g$ on $B_1(S)$. </p>\n\n<p>Now,\n$$\n\\mathrm{d}\\omega^2 = -\\theta^2_j\\wedge\\omega^j \n= -g\\,\\omega^2\\wedge\\omega^1 - \\theta^2_2\\wedge\\omega^2\n= -(\\theta^2_2 - g\\,\\omega^1)\\wedge\\omega^2.\n$$ \nThus, $\\omega^2$ is an integrable $1$-form, and, because it is semi-basic\nfor the submersion $x:B_1(S)\\to S\\subset\\mathbb{R}^4$, it follows that $\\omega^2$ is a multiple of the $x$-pullback of a (nonvanishing) $1$-form on $S$. Thus, $S$ is foliated by (connected) curves whose $x$-preimages in $B_1(S)$ are codimension $1$ integral submanifolds of $\\omega^2$.</p>\n\n<p>I claim that these curves in $S$ are, in fact, lines in $\\mathbb{R}^4$. To see this, note that, when one pulls back to a leaf of $\\omega^2=0$ in $B_1(S)$, one has $\\theta^2_1 = \\theta^3_1=\\theta^4_1= 0$ as well, so one has\n$$\n\\mathrm{d}x = e_1\\,\\omega^1\\qquad\\text{and}\\qquad \\mathrm{d}e_1 = e_1\\,\\theta^1_1\\,.\n$$\nIn particular, the direction of $e_1$ is fixed on this leaf, and this is the tangent direction of the mapping $x$ restricted to this leaf. Hence the $x$-image of this leaf is an open interval in a line in $\\mathbb{R}^4$. Thus, the surface is ruled, as claimed.<p></p>\n\n<p>There still remains the question of how one could 'generate' these surfaces, at least locally. I claim that, since these surfaces are, in fact, $1$-parameter families of lines, one should think of them as curves in the space of lines that satisfy some differential equations. Here is how one can think of this system of equations:<p></p>\n\n<p>Let $\\mathcal{I}$ denote the Pfaffian system of rank $5$ that is generated by the five linearly independent $1$-forms\n$$\n\\omega^3,\\ \\omega^4,\\ \\theta^4_1\\,,\\ \\theta^4_2\\,,\\ \\theta^3_1 \n$$\n(these are the $1$-forms that vanish when pulled back to $B_1(S)$ when $S\\subset\\mathbb{R}^4$ is a surface satisfying our hypotheses). By the structure equations, \n$$\n\\left.\n\\begin{aligned}\n\\mathrm{d}\\omega^4 &amp;\\equiv 0\\\\\n\\mathrm{d}\\theta^4_1 &amp;\\equiv 0\\\\\n\\mathrm{d}\\omega^3 &amp;\\equiv -\\theta^3_2\\wedge\\omega^2\\\\\n\\mathrm{d}\\theta^3_1 &amp;\\equiv -\\theta^3_2\\wedge\\theta^2_1\\\\\n\\mathrm{d}\\theta^4_2 &amp;\\equiv \\phantom{-}\\theta^3_2\\wedge\\theta^4_3\\\\\n\\end{aligned}\\ \\right\\} \\mathrm{modulo}\\ \\mathcal{I}\n$$\nIt follows that $\\mathcal{J}$, the Cartan system of $\\mathcal{I}$, \nhas rank $9$ and is spanned by the nine $1$-forms\n$$\n\\omega^3,\\ \\omega^4,\\ \\theta^4_1\\,,\\ \\theta^4_2\\,,\\ \\theta^3_1\\,,\\ \n\\omega^2,\\ \\theta^2_1\\,,\\ \\theta^3_2\\,,\\ \\theta^4_3\n$$\nIn fact, $\\mathcal{J}$ is easily seen to be the bundle of $1$-forms \non $\\mathbb{R}^4\\times\\mathrm{GL}(4,\\mathbb{R})$ that is semibasic for the projection to the $9$-manifold $F$, where the projection sends $(p; v_i)$\nto the element of $F$ described by\n$$\n\\bigl (\\lambda(p,v_1),\\ [v_1\\wedge v_2], [v_1\\wedge v_2\\wedge v_3]\\ \\bigr)\n$$\nand where $\\lambda(p,v_1)$ is the line through $p$ in the direction $v_1$.</p>\n\n<p>In particular $\\mathcal{I}$ is the pullback of a well-defined Pfaffian system of rank $5$ on $F$, whose annihilator is the $4$-plane field $D\\subset TF$. Moreover, it is easy to show that the three rank $6$ Pfaffian systems generated by adjoining any one of $\\omega^2$, $\\theta^3_2$, or $\\theta^4_3$ to $\\mathcal{I}$ are themselves pullbacks of rank $6$ Pfaffian systems on $F$ whose annihilators in $TF$ are each $3$-plane subbundles of $D$. <p>\nBy its very construction, the projection of $B_1(S)$ into $F$ is a curve that is tangent to $D$ and not tangent to any of these three $3$-plane subbundles of $D$.</p>\n\n<p>Conversely, if $\\gamma\\subset F$ is any integral curve of $D$ that is not tangent to any of the three $3$-plane subbundles of $D$, its preimage in $\\mathbb{R}^4\\times\\mathrm{GL}(4,\\mathbb{R})$ is a submanifold of the form $B_1(S)$ for a surface $S$ satisfying our conditions, in fact, the surface swept out by the union of the lines represented by $\\lambda(\\gamma)$, where $\\lambda:F\\to \\Lambda$ is the obvious map to the lines.</p>\n\n<p>It is a standard fact that the curves tangent to a 4-plane field in a smooth manifold are locally described (up to reparametrization) by prescribing $3$ functions of one variable.</p>\n" } ]
Question: A variant of the Monge-Cayley-Salmon theorem? Answer 1: <p>Setting aside the assumption that $\phi$ be a polynomial mapping for the moment (however, see below for a construction of a large family of polynomial solutions), if one makes the 'nondegeneracy' assumptions</p> <ol> <li>$\mathrm{dim}\ \mathrm{span}\bigl( \phi_s(s,t), \phi_t(s,t)\bigr) =2 $,</li> <li>$\mathrm{dim}\ \mathrm{span}\bigl( \phi_s(s,t), \phi_t(s,t), \phi_{ss}(s,t), \phi_{st}(s,t), \phi_{tt}(s,t) \bigr) = 3$ for all $(s,t)\in[0,1]^2$, and </li> <li>the subspace $W(s,t) =\mathrm{span}\bigl( \phi_s(s,t), \phi_t(s,t), \phi_{ss}(s,t), \phi_{st}(s,t), \phi_{tt}(s,t) \bigr)\subset\mathbb{R}^4$ is not constant, in the sense that $W:[0,1]^2\to \mathrm{Gr}_3(\mathbb{R}^4)\simeq\mathbb{RP}^3$ has nonvanishing differential,</li> </ol> <p>then one can show that the surface $\phi\bigl([0,1]^2\bigr)\subset\mathbb{R}^4$ is ruled (and does not lie in an affine $3$-space). </p> <p>Such surfaces locally depend on three arbitrary functions of one variable in Cartan's sense. One way of describing them is this: Let $\Lambda$ be the space of (affine) lines in $\mathbb{R}^4$, a manifold of dimension $6$. Consider the ($9$-dimensional) bundle $\pi:F\to\Lambda$ whose fiber over $\lambda\in\Lambda$ is the flag variety of the $3$-dimensional vector space $\mathbb{R}^4/\lambda'$, where $\lambda'\subset\mathbb{R}^4$ is the linear subspace parallel to $\lambda$. Then there exists a smooth $4$-plane field $D\subset TF$ such that, if $\gamma\subset F$ is a generic curve tangent to $D$, then regarding $\gamma$ as a $1$-parameter family of affine lines via $\pi(\gamma)\subset \Lambda$, the union of these lines is a surface satisfying the above assumptions. (Here 'generic' means that the tangents to $\gamma$ do not lie in triple of hyperplanes in $D$.)</p> <p>It is not hard to write down polynomial solutions: For example, if $f:\mathbb{R}\to\mathbb{R}^4$ is a polynomial curve that satisfies $f'(s)\wedge f''(s)\wedge f'''(s) \wedge f''''(s)\not = 0$, then the mapping $\phi:\mathbb{R}^2\to\mathbb{R}^4$ given by $$ \phi(s,t) = f(s) + f'(s)\, t $$ (which parametrizes the 'tangential development' of the curve $f$) satisfies these conditions when $t\not=0$. (By replacing $t$ by $t{+}1$, say, one could arrange that $\phi$ be an immersion on all of $[0,1]^2$.)</p> <p>The following analysis is a more-or-less standard approach to verifying the above description using the so-called 'moving frame'. (I'm sure that the result itself is classical in some sense, though I don't know offhand where to look in the literature to find it.)</p> <p>The three conditions listed above are actually independent of the choice of $st$-coordinates on the surface in $\mathbb{R}^4$ and so can be regarded as conditions on a surface $S\subset\mathbb{R}^4$ that, for local analysis purposes, can be taken to be smoothly embedded.</p> <p>Let $B_0(S)$ denote the space of quintuples $(p;v_1,v_2,v_3,v_4)$ where $p$ lies in $S$ and the quadruple $(v_1,v_2,v_3,v_4)$ is a basis of $\mathbb{R}^4$ such that $(v_1,v_2)$ is a basis of $T_pS$ while $(v_1,v_2,v_3)$ is a basis of the $3$-dimensional subspace $W_pS\subset \mathbb{R}^4$. Then $B_0$ is a smooth submanifold of the product $\mathbb{R}^4\times\mathrm{GL}(4,\mathbb{R})$ that has dimension $2 + 4 + 3 + 4 = 13$. <p> It is useful to define $\mathbb{R}^4$-valued functions $x,e_i: \mathbb{R}^4\times\mathrm{GL}(4,\mathbb{R})\to\mathbb{R}^4$ such that $x(p;v_1,v_2,v_3,v_4) = p$ while $e_i(p;v_1,v_2,v_3,v_4) = v_i$. Then there exist unique (linearly independent) $1$-forms $\omega^i$ and $\theta^i_j$ on $\mathbb{R}^4\times\mathrm{GL}(4,\mathbb{R})$ such that (assuming the usual summation convention on repeated indices) the following <em>structure equations</em> hold: $$ \mathrm{d}x = e_i \omega^i \qquad\text{and}\qquad \mathrm{d}e_i = e_j\,\theta^j_i\,.\tag1 $$ </p> <p>$$ \mathrm{d}\omega^i = -\theta^i_j\wedge\omega^j \qquad\text{and}\qquad \mathrm{d}\theta^i_j = -\theta^i_k\wedge\theta^k_j\,.\tag2 $$</p> <p>Now, pull back these functions and $1$-forms to $B_0(S)$ (but, as is customary, not notate the pullback). The definition of $B_0(S)$ and the assumptions on $S$ imply that $$ \mathrm{d}x\wedge e_1\wedge e_2 = \mathrm{d}e_1\wedge e_1\wedge e_2\wedge e_3 = \mathrm{d}e_2\wedge e_1\wedge e_2\wedge e_3 = 0 $$ while the two expressions $$ \bigl(\,\mathrm{d}e_1\wedge e_1\wedge e_2,\ \mathrm{d}e_2\wedge e_1\wedge e_2\bigr) \qquad\text{and}\qquad \mathrm{d}e_3\wedge e_1\wedge e_2\wedge e_3 $$ are nowhere vanishing. Using the above structure equations, these imply the relations $$ \omega^3 = \omega^4 = \theta^4_1 = \theta^4_2 = 0,\tag3 $$ while, because of the three assumptions, $\omega^1\wedge\omega^2$ is nonvanishing, the pair $(\theta^3_1,\theta^3_2)$ do not simultaneously vanish, and $\theta^4_3$ is nonvanishing.</p> <p>Meanwhile, the structure equations yield $$ 0 = \mathrm{d}\omega^3 = -\theta^3_1\wedge\omega^1 -\theta^3_2\wedge\omega^2, $$ so there must exist functions $h_{ij}=h_{ji}$ for $1\le i,j,\le 2$, not all simultaneously vanishing, such that $\theta^3_i = h_{ij}\omega^j$. The quadratic form $h = h_{ij}\,\omega^i\omega^j$ is then nonvanishing and well-defined up to multiples on the surface $S$. Moreover, for $i = 1$ or $2$, $$ 0 = \mathrm{d}\theta^4_i = -\theta^4_k\wedge\theta^k_i = -\theta^4_3\wedge\theta^3_i\,. $$ Thus, since $\theta^4_3$ is nonvanishing, it follows that $\theta^3_1$ and $\theta^3_2$ are multiples of $\theta^4_3$. In particular, $\theta^3_1\wedge\theta^3_2$ vanishes identically, so $h_{11}h_{22}-{h_{12}}^2$ vanishes identically. Thus, the quadratic form $h$ has rank $1$.</p> <p>Let $B_1(S)\subset B_2(S)$ denote the submanifold defined by $h_{11} = h_{12}=0$. It is a smooth submanifold of $B_0(S)$ of dimension $11$, and when all the forms and functions are pulled back to $B_1(S)$, we have $\theta^3_1 = 0$ while $\theta^3_2 = h_{22}\,\omega^2$. In particular, it now follows that $\theta^4_3$ is also a multiple of $\omega^2$, say $\theta^4_3 = f\,\omega^2$ for some $f$ (which is nonvanishing).</p> <p>Moreover, $$ 0 = \mathrm{d}\theta^3_1 = -\theta^3_k\wedge\theta^k_1 = -\theta^3_2\wedge \theta^2_1 = -h_{22}\,\omega^2\wedge\theta^2_1, $$ so it follows (since $h_{22}$ is nonvanishing) that $\theta^2_1 = g\,\omega^2$ for some function $g$ on $B_1(S)$. </p> <p>Now, $$ \mathrm{d}\omega^2 = -\theta^2_j\wedge\omega^j = -g\,\omega^2\wedge\omega^1 - \theta^2_2\wedge\omega^2 = -(\theta^2_2 - g\,\omega^1)\wedge\omega^2. $$ Thus, $\omega^2$ is an integrable $1$-form, and, because it is semi-basic for the submersion $x:B_1(S)\to S\subset\mathbb{R}^4$, it follows that $\omega^2$ is a multiple of the $x$-pullback of a (nonvanishing) $1$-form on $S$. Thus, $S$ is foliated by (connected) curves whose $x$-preimages in $B_1(S)$ are codimension $1$ integral submanifolds of $\omega^2$.</p> <p>I claim that these curves in $S$ are, in fact, lines in $\mathbb{R}^4$. To see this, note that, when one pulls back to a leaf of $\omega^2=0$ in $B_1(S)$, one has $\theta^2_1 = \theta^3_1=\theta^4_1= 0$ as well, so one has $$ \mathrm{d}x = e_1\,\omega^1\qquad\text{and}\qquad \mathrm{d}e_1 = e_1\,\theta^1_1\,. $$ In particular, the direction of $e_1$ is fixed on this leaf, and this is the tangent direction of the mapping $x$ restricted to this leaf. Hence the $x$-image of this leaf is an open interval in a line in $\mathbb{R}^4$. Thus, the surface is ruled, as claimed.<p></p> <p>There still remains the question of how one could 'generate' these surfaces, at least locally. I claim that, since these surfaces are, in fact, $1$-parameter families of lines, one should think of them as curves in the space of lines that satisfy some differential equations. Here is how one can think of this system of equations:<p></p> <p>Let $\mathcal{I}$ denote the Pfaffian system of rank $5$ that is generated by the five linearly independent $1$-forms $$ \omega^3,\ \omega^4,\ \theta^4_1\,,\ \theta^4_2\,,\ \theta^3_1 $$ (these are the $1$-forms that vanish when pulled back to $B_1(S)$ when $S\subset\mathbb{R}^4$ is a surface satisfying our hypotheses). By the structure equations, $$ \left. \begin{aligned} \mathrm{d}\omega^4 &amp;\equiv 0\\ \mathrm{d}\theta^4_1 &amp;\equiv 0\\ \mathrm{d}\omega^3 &amp;\equiv -\theta^3_2\wedge\omega^2\\ \mathrm{d}\theta^3_1 &amp;\equiv -\theta^3_2\wedge\theta^2_1\\ \mathrm{d}\theta^4_2 &amp;\equiv \phantom{-}\theta^3_2\wedge\theta^4_3\\ \end{aligned}\ \right\} \mathrm{modulo}\ \mathcal{I} $$ It follows that $\mathcal{J}$, the Cartan system of $\mathcal{I}$, has rank $9$ and is spanned by the nine $1$-forms $$ \omega^3,\ \omega^4,\ \theta^4_1\,,\ \theta^4_2\,,\ \theta^3_1\,,\ \omega^2,\ \theta^2_1\,,\ \theta^3_2\,,\ \theta^4_3 $$ In fact, $\mathcal{J}$ is easily seen to be the bundle of $1$-forms on $\mathbb{R}^4\times\mathrm{GL}(4,\mathbb{R})$ that is semibasic for the projection to the $9$-manifold $F$, where the projection sends $(p; v_i)$ to the element of $F$ described by $$ \bigl (\lambda(p,v_1),\ [v_1\wedge v_2], [v_1\wedge v_2\wedge v_3]\ \bigr) $$ and where $\lambda(p,v_1)$ is the line through $p$ in the direction $v_1$.</p> <p>In particular $\mathcal{I}$ is the pullback of a well-defined Pfaffian system of rank $5$ on $F$, whose annihilator is the $4$-plane field $D\subset TF$. Moreover, it is easy to show that the three rank $6$ Pfaffian systems generated by adjoining any one of $\omega^2$, $\theta^3_2$, or $\theta^4_3$ to $\mathcal{I}$ are themselves pullbacks of rank $6$ Pfaffian systems on $F$ whose annihilators in $TF$ are each $3$-plane subbundles of $D$. <p> By its very construction, the projection of $B_1(S)$ into $F$ is a curve that is tangent to $D$ and not tangent to any of these three $3$-plane subbundles of $D$.</p> <p>Conversely, if $\gamma\subset F$ is any integral curve of $D$ that is not tangent to any of the three $3$-plane subbundles of $D$, its preimage in $\mathbb{R}^4\times\mathrm{GL}(4,\mathbb{R})$ is a submanifold of the form $B_1(S)$ for a surface $S$ satisfying our conditions, in fact, the surface swept out by the union of the lines represented by $\lambda(\gamma)$, where $\lambda:F\to \Lambda$ is the obvious map to the lines.</p> <p>It is a standard fact that the curves tangent to a 4-plane field in a smooth manifold are locally described (up to reparametrization) by prescribing $3$ functions of one variable.</p> Answer 2: <p>Setting aside the assumption that $\phi$ be a polynomial mapping for the moment (however, see below for a construction of a large family of polynomial solutions), if one makes the 'nondegeneracy' assumptions</p> <ol> <li>$\mathrm{dim}\ \mathrm{span}\bigl( \phi_s(s,t), \phi_t(s,t)\bigr) =2 $,</li> <li>$\mathrm{dim}\ \mathrm{span}\bigl( \phi_s(s,t), \phi_t(s,t), \phi_{ss}(s,t), \phi_{st}(s,t), \phi_{tt}(s,t) \bigr) = 3$ for all $(s,t)\in[0,1]^2$, and </li> <li>the subspace $W(s,t) =\mathrm{span}\bigl( \phi_s(s,t), \phi_t(s,t), \phi_{ss}(s,t), \phi_{st}(s,t), \phi_{tt}(s,t) \bigr)\subset\mathbb{R}^4$ is not constant, in the sense that $W:[0,1]^2\to \mathrm{Gr}_3(\mathbb{R}^4)\simeq\mathbb{RP}^3$ has nonvanishing differential,</li> </ol> <p>then one can show that the surface $\phi\bigl([0,1]^2\bigr)\subset\mathbb{R}^4$ is ruled (and does not lie in an affine $3$-space). </p> <p>Such surfaces locally depend on three arbitrary functions of one variable in Cartan's sense. One way of describing them is this: Let $\Lambda$ be the space of (affine) lines in $\mathbb{R}^4$, a manifold of dimension $6$. Consider the ($9$-dimensional) bundle $\pi:F\to\Lambda$ whose fiber over $\lambda\in\Lambda$ is the flag variety of the $3$-dimensional vector space $\mathbb{R}^4/\lambda'$, where $\lambda'\subset\mathbb{R}^4$ is the linear subspace parallel to $\lambda$. Then there exists a smooth $4$-plane field $D\subset TF$ such that, if $\gamma\subset F$ is a generic curve tangent to $D$, then regarding $\gamma$ as a $1$-parameter family of affine lines via $\pi(\gamma)\subset \Lambda$, the union of these lines is a surface satisfying the above assumptions. (Here 'generic' means that the tangents to $\gamma$ do not lie in triple of hyperplanes in $D$.)</p> <p>It is not hard to write down polynomial solutions: For example, if $f:\mathbb{R}\to\mathbb{R}^4$ is a polynomial curve that satisfies $f'(s)\wedge f''(s)\wedge f'''(s) \wedge f''''(s)\not = 0$, then the mapping $\phi:\mathbb{R}^2\to\mathbb{R}^4$ given by $$ \phi(s,t) = f(s) + f'(s)\, t $$ (which parametrizes the 'tangential development' of the curve $f$) satisfies these conditions when $t\not=0$. (By replacing $t$ by $t{+}1$, say, one could arrange that $\phi$ be an immersion on all of $[0,1]^2$.)</p> <p>The following analysis is a more-or-less standard approach to verifying the above description using the so-called 'moving frame'. (I'm sure that the result itself is classical in some sense, though I don't know offhand where to look in the literature to find it.)</p> <p>The three conditions listed above are actually independent of the choice of $st$-coordinates on the surface in $\mathbb{R}^4$ and so can be regarded as conditions on a surface $S\subset\mathbb{R}^4$ that, for local analysis purposes, can be taken to be smoothly embedded.</p> <p>Let $B_0(S)$ denote the space of quintuples $(p;v_1,v_2,v_3,v_4)$ where $p$ lies in $S$ and the quadruple $(v_1,v_2,v_3,v_4)$ is a basis of $\mathbb{R}^4$ such that $(v_1,v_2)$ is a basis of $T_pS$ while $(v_1,v_2,v_3)$ is a basis of the $3$-dimensional subspace $W_pS\subset \mathbb{R}^4$. Then $B_0$ is a smooth submanifold of the product $\mathbb{R}^4\times\mathrm{GL}(4,\mathbb{R})$ that has dimension $2 + 4 + 3 + 4 = 13$. <p> It is useful to define $\mathbb{R}^4$-valued functions $x,e_i: \mathbb{R}^4\times\mathrm{GL}(4,\mathbb{R})\to\mathbb{R}^4$ such that $x(p;v_1,v_2,v_3,v_4) = p$ while $e_i(p;v_1,v_2,v_3,v_4) = v_i$. Then there exist unique (linearly independent) $1$-forms $\omega^i$ and $\theta^i_j$ on $\mathbb{R}^4\times\mathrm{GL}(4,\mathbb{R})$ such that (assuming the usual summation convention on repeated indices) the following <em>structure equations</em> hold: $$ \mathrm{d}x = e_i \omega^i \qquad\text{and}\qquad \mathrm{d}e_i = e_j\,\theta^j_i\,.\tag1 $$ </p> <p>$$ \mathrm{d}\omega^i = -\theta^i_j\wedge\omega^j \qquad\text{and}\qquad \mathrm{d}\theta^i_j = -\theta^i_k\wedge\theta^k_j\,.\tag2 $$</p> <p>Now, pull back these functions and $1$-forms to $B_0(S)$ (but, as is customary, not notate the pullback). The definition of $B_0(S)$ and the assumptions on $S$ imply that $$ \mathrm{d}x\wedge e_1\wedge e_2 = \mathrm{d}e_1\wedge e_1\wedge e_2\wedge e_3 = \mathrm{d}e_2\wedge e_1\wedge e_2\wedge e_3 = 0 $$ while the two expressions $$ \bigl(\,\mathrm{d}e_1\wedge e_1\wedge e_2,\ \mathrm{d}e_2\wedge e_1\wedge e_2\bigr) \qquad\text{and}\qquad \mathrm{d}e_3\wedge e_1\wedge e_2\wedge e_3 $$ are nowhere vanishing. Using the above structure equations, these imply the relations $$ \omega^3 = \omega^4 = \theta^4_1 = \theta^4_2 = 0,\tag3 $$ while, because of the three assumptions, $\omega^1\wedge\omega^2$ is nonvanishing, the pair $(\theta^3_1,\theta^3_2)$ do not simultaneously vanish, and $\theta^4_3$ is nonvanishing.</p> <p>Meanwhile, the structure equations yield $$ 0 = \mathrm{d}\omega^3 = -\theta^3_1\wedge\omega^1 -\theta^3_2\wedge\omega^2, $$ so there must exist functions $h_{ij}=h_{ji}$ for $1\le i,j,\le 2$, not all simultaneously vanishing, such that $\theta^3_i = h_{ij}\omega^j$. The quadratic form $h = h_{ij}\,\omega^i\omega^j$ is then nonvanishing and well-defined up to multiples on the surface $S$. Moreover, for $i = 1$ or $2$, $$ 0 = \mathrm{d}\theta^4_i = -\theta^4_k\wedge\theta^k_i = -\theta^4_3\wedge\theta^3_i\,. $$ Thus, since $\theta^4_3$ is nonvanishing, it follows that $\theta^3_1$ and $\theta^3_2$ are multiples of $\theta^4_3$. In particular, $\theta^3_1\wedge\theta^3_2$ vanishes identically, so $h_{11}h_{22}-{h_{12}}^2$ vanishes identically. Thus, the quadratic form $h$ has rank $1$.</p> <p>Let $B_1(S)\subset B_2(S)$ denote the submanifold defined by $h_{11} = h_{12}=0$. It is a smooth submanifold of $B_0(S)$ of dimension $11$, and when all the forms and functions are pulled back to $B_1(S)$, we have $\theta^3_1 = 0$ while $\theta^3_2 = h_{22}\,\omega^2$. In particular, it now follows that $\theta^4_3$ is also a multiple of $\omega^2$, say $\theta^4_3 = f\,\omega^2$ for some $f$ (which is nonvanishing).</p> <p>Moreover, $$ 0 = \mathrm{d}\theta^3_1 = -\theta^3_k\wedge\theta^k_1 = -\theta^3_2\wedge \theta^2_1 = -h_{22}\,\omega^2\wedge\theta^2_1, $$ so it follows (since $h_{22}$ is nonvanishing) that $\theta^2_1 = g\,\omega^2$ for some function $g$ on $B_1(S)$. </p> <p>Now, $$ \mathrm{d}\omega^2 = -\theta^2_j\wedge\omega^j = -g\,\omega^2\wedge\omega^1 - \theta^2_2\wedge\omega^2 = -(\theta^2_2 - g\,\omega^1)\wedge\omega^2. $$ Thus, $\omega^2$ is an integrable $1$-form, and, because it is semi-basic for the submersion $x:B_1(S)\to S\subset\mathbb{R}^4$, it follows that $\omega^2$ is a multiple of the $x$-pullback of a (nonvanishing) $1$-form on $S$. Thus, $S$ is foliated by (connected) curves whose $x$-preimages in $B_1(S)$ are codimension $1$ integral submanifolds of $\omega^2$.</p> <p>I claim that these curves in $S$ are, in fact, lines in $\mathbb{R}^4$. To see this, note that, when one pulls back to a leaf of $\omega^2=0$ in $B_1(S)$, one has $\theta^2_1 = \theta^3_1=\theta^4_1= 0$ as well, so one has $$ \mathrm{d}x = e_1\,\omega^1\qquad\text{and}\qquad \mathrm{d}e_1 = e_1\,\theta^1_1\,. $$ In particular, the direction of $e_1$ is fixed on this leaf, and this is the tangent direction of the mapping $x$ restricted to this leaf. Hence the $x$-image of this leaf is an open interval in a line in $\mathbb{R}^4$. Thus, the surface is ruled, as claimed.<p></p> <p>There still remains the question of how one could 'generate' these surfaces, at least locally. I claim that, since these surfaces are, in fact, $1$-parameter families of lines, one should think of them as curves in the space of lines that satisfy some differential equations. Here is how one can think of this system of equations:<p></p> <p>Let $\mathcal{I}$ denote the Pfaffian system of rank $5$ that is generated by the five linearly independent $1$-forms $$ \omega^3,\ \omega^4,\ \theta^4_1\,,\ \theta^4_2\,,\ \theta^3_1 $$ (these are the $1$-forms that vanish when pulled back to $B_1(S)$ when $S\subset\mathbb{R}^4$ is a surface satisfying our hypotheses). By the structure equations, $$ \left. \begin{aligned} \mathrm{d}\omega^4 &amp;\equiv 0\\ \mathrm{d}\theta^4_1 &amp;\equiv 0\\ \mathrm{d}\omega^3 &amp;\equiv -\theta^3_2\wedge\omega^2\\ \mathrm{d}\theta^3_1 &amp;\equiv -\theta^3_2\wedge\theta^2_1\\ \mathrm{d}\theta^4_2 &amp;\equiv \phantom{-}\theta^3_2\wedge\theta^4_3\\ \end{aligned}\ \right\} \mathrm{modulo}\ \mathcal{I} $$ It follows that $\mathcal{J}$, the Cartan system of $\mathcal{I}$, has rank $9$ and is spanned by the nine $1$-forms $$ \omega^3,\ \omega^4,\ \theta^4_1\,,\ \theta^4_2\,,\ \theta^3_1\,,\ \omega^2,\ \theta^2_1\,,\ \theta^3_2\,,\ \theta^4_3 $$ In fact, $\mathcal{J}$ is easily seen to be the bundle of $1$-forms on $\mathbb{R}^4\times\mathrm{GL}(4,\mathbb{R})$ that is semibasic for the projection to the $9$-manifold $F$, where the projection sends $(p; v_i)$ to the element of $F$ described by $$ \bigl (\lambda(p,v_1),\ [v_1\wedge v_2], [v_1\wedge v_2\wedge v_3]\ \bigr) $$ and where $\lambda(p,v_1)$ is the line through $p$ in the direction $v_1$.</p> <p>In particular $\mathcal{I}$ is the pullback of a well-defined Pfaffian system of rank $5$ on $F$, whose annihilator is the $4$-plane field $D\subset TF$. Moreover, it is easy to show that the three rank $6$ Pfaffian systems generated by adjoining any one of $\omega^2$, $\theta^3_2$, or $\theta^4_3$ to $\mathcal{I}$ are themselves pullbacks of rank $6$ Pfaffian systems on $F$ whose annihilators in $TF$ are each $3$-plane subbundles of $D$. <p> By its very construction, the projection of $B_1(S)$ into $F$ is a curve that is tangent to $D$ and not tangent to any of these three $3$-plane subbundles of $D$.</p> <p>Conversely, if $\gamma\subset F$ is any integral curve of $D$ that is not tangent to any of the three $3$-plane subbundles of $D$, its preimage in $\mathbb{R}^4\times\mathrm{GL}(4,\mathbb{R})$ is a submanifold of the form $B_1(S)$ for a surface $S$ satisfying our conditions, in fact, the surface swept out by the union of the lines represented by $\lambda(\gamma)$, where $\lambda:F\to \Lambda$ is the obvious map to the lines.</p> <p>It is a standard fact that the curves tangent to a 4-plane field in a smooth manifold are locally described (up to reparametrization) by prescribing $3$ functions of one variable.</p> Answer 3: <p>Setting aside the assumption that $\phi$ be a polynomial mapping for the moment (however, see below for a construction of a large family of polynomial solutions), if one makes the 'nondegeneracy' assumptions</p> <ol> <li>$\mathrm{dim}\ \mathrm{span}\bigl( \phi_s(s,t), \phi_t(s,t)\bigr) =2 $,</li> <li>$\mathrm{dim}\ \mathrm{span}\bigl( \phi_s(s,t), \phi_t(s,t), \phi_{ss}(s,t), \phi_{st}(s,t), \phi_{tt}(s,t) \bigr) = 3$ for all $(s,t)\in[0,1]^2$, and </li> <li>the subspace $W(s,t) =\mathrm{span}\bigl( \phi_s(s,t), \phi_t(s,t), \phi_{ss}(s,t), \phi_{st}(s,t), \phi_{tt}(s,t) \bigr)\subset\mathbb{R}^4$ is not constant, in the sense that $W:[0,1]^2\to \mathrm{Gr}_3(\mathbb{R}^4)\simeq\mathbb{RP}^3$ has nonvanishing differential,</li> </ol> <p>then one can show that the surface $\phi\bigl([0,1]^2\bigr)\subset\mathbb{R}^4$ is ruled (and does not lie in an affine $3$-space). </p> <p>Such surfaces locally depend on three arbitrary functions of one variable in Cartan's sense. One way of describing them is this: Let $\Lambda$ be the space of (affine) lines in $\mathbb{R}^4$, a manifold of dimension $6$. Consider the ($9$-dimensional) bundle $\pi:F\to\Lambda$ whose fiber over $\lambda\in\Lambda$ is the flag variety of the $3$-dimensional vector space $\mathbb{R}^4/\lambda'$, where $\lambda'\subset\mathbb{R}^4$ is the linear subspace parallel to $\lambda$. Then there exists a smooth $4$-plane field $D\subset TF$ such that, if $\gamma\subset F$ is a generic curve tangent to $D$, then regarding $\gamma$ as a $1$-parameter family of affine lines via $\pi(\gamma)\subset \Lambda$, the union of these lines is a surface satisfying the above assumptions. (Here 'generic' means that the tangents to $\gamma$ do not lie in triple of hyperplanes in $D$.)</p> <p>It is not hard to write down polynomial solutions: For example, if $f:\mathbb{R}\to\mathbb{R}^4$ is a polynomial curve that satisfies $f'(s)\wedge f''(s)\wedge f'''(s) \wedge f''''(s)\not = 0$, then the mapping $\phi:\mathbb{R}^2\to\mathbb{R}^4$ given by $$ \phi(s,t) = f(s) + f'(s)\, t $$ (which parametrizes the 'tangential development' of the curve $f$) satisfies these conditions when $t\not=0$. (By replacing $t$ by $t{+}1$, say, one could arrange that $\phi$ be an immersion on all of $[0,1]^2$.)</p> <p>The following analysis is a more-or-less standard approach to verifying the above description using the so-called 'moving frame'. (I'm sure that the result itself is classical in some sense, though I don't know offhand where to look in the literature to find it.)</p> <p>The three conditions listed above are actually independent of the choice of $st$-coordinates on the surface in $\mathbb{R}^4$ and so can be regarded as conditions on a surface $S\subset\mathbb{R}^4$ that, for local analysis purposes, can be taken to be smoothly embedded.</p> <p>Let $B_0(S)$ denote the space of quintuples $(p;v_1,v_2,v_3,v_4)$ where $p$ lies in $S$ and the quadruple $(v_1,v_2,v_3,v_4)$ is a basis of $\mathbb{R}^4$ such that $(v_1,v_2)$ is a basis of $T_pS$ while $(v_1,v_2,v_3)$ is a basis of the $3$-dimensional subspace $W_pS\subset \mathbb{R}^4$. Then $B_0$ is a smooth submanifold of the product $\mathbb{R}^4\times\mathrm{GL}(4,\mathbb{R})$ that has dimension $2 + 4 + 3 + 4 = 13$. <p> It is useful to define $\mathbb{R}^4$-valued functions $x,e_i: \mathbb{R}^4\times\mathrm{GL}(4,\mathbb{R})\to\mathbb{R}^4$ such that $x(p;v_1,v_2,v_3,v_4) = p$ while $e_i(p;v_1,v_2,v_3,v_4) = v_i$. Then there exist unique (linearly independent) $1$-forms $\omega^i$ and $\theta^i_j$ on $\mathbb{R}^4\times\mathrm{GL}(4,\mathbb{R})$ such that (assuming the usual summation convention on repeated indices) the following <em>structure equations</em> hold: $$ \mathrm{d}x = e_i \omega^i \qquad\text{and}\qquad \mathrm{d}e_i = e_j\,\theta^j_i\,.\tag1 $$ </p> <p>$$ \mathrm{d}\omega^i = -\theta^i_j\wedge\omega^j \qquad\text{and}\qquad \mathrm{d}\theta^i_j = -\theta^i_k\wedge\theta^k_j\,.\tag2 $$</p> <p>Now, pull back these functions and $1$-forms to $B_0(S)$ (but, as is customary, not notate the pullback). The definition of $B_0(S)$ and the assumptions on $S$ imply that $$ \mathrm{d}x\wedge e_1\wedge e_2 = \mathrm{d}e_1\wedge e_1\wedge e_2\wedge e_3 = \mathrm{d}e_2\wedge e_1\wedge e_2\wedge e_3 = 0 $$ while the two expressions $$ \bigl(\,\mathrm{d}e_1\wedge e_1\wedge e_2,\ \mathrm{d}e_2\wedge e_1\wedge e_2\bigr) \qquad\text{and}\qquad \mathrm{d}e_3\wedge e_1\wedge e_2\wedge e_3 $$ are nowhere vanishing. Using the above structure equations, these imply the relations $$ \omega^3 = \omega^4 = \theta^4_1 = \theta^4_2 = 0,\tag3 $$ while, because of the three assumptions, $\omega^1\wedge\omega^2$ is nonvanishing, the pair $(\theta^3_1,\theta^3_2)$ do not simultaneously vanish, and $\theta^4_3$ is nonvanishing.</p> <p>Meanwhile, the structure equations yield $$ 0 = \mathrm{d}\omega^3 = -\theta^3_1\wedge\omega^1 -\theta^3_2\wedge\omega^2, $$ so there must exist functions $h_{ij}=h_{ji}$ for $1\le i,j,\le 2$, not all simultaneously vanishing, such that $\theta^3_i = h_{ij}\omega^j$. The quadratic form $h = h_{ij}\,\omega^i\omega^j$ is then nonvanishing and well-defined up to multiples on the surface $S$. Moreover, for $i = 1$ or $2$, $$ 0 = \mathrm{d}\theta^4_i = -\theta^4_k\wedge\theta^k_i = -\theta^4_3\wedge\theta^3_i\,. $$ Thus, since $\theta^4_3$ is nonvanishing, it follows that $\theta^3_1$ and $\theta^3_2$ are multiples of $\theta^4_3$. In particular, $\theta^3_1\wedge\theta^3_2$ vanishes identically, so $h_{11}h_{22}-{h_{12}}^2$ vanishes identically. Thus, the quadratic form $h$ has rank $1$.</p> <p>Let $B_1(S)\subset B_2(S)$ denote the submanifold defined by $h_{11} = h_{12}=0$. It is a smooth submanifold of $B_0(S)$ of dimension $11$, and when all the forms and functions are pulled back to $B_1(S)$, we have $\theta^3_1 = 0$ while $\theta^3_2 = h_{22}\,\omega^2$. In particular, it now follows that $\theta^4_3$ is also a multiple of $\omega^2$, say $\theta^4_3 = f\,\omega^2$ for some $f$ (which is nonvanishing).</p> <p>Moreover, $$ 0 = \mathrm{d}\theta^3_1 = -\theta^3_k\wedge\theta^k_1 = -\theta^3_2\wedge \theta^2_1 = -h_{22}\,\omega^2\wedge\theta^2_1, $$ so it follows (since $h_{22}$ is nonvanishing) that $\theta^2_1 = g\,\omega^2$ for some function $g$ on $B_1(S)$. </p> <p>Now, $$ \mathrm{d}\omega^2 = -\theta^2_j\wedge\omega^j = -g\,\omega^2\wedge\omega^1 - \theta^2_2\wedge\omega^2 = -(\theta^2_2 - g\,\omega^1)\wedge\omega^2. $$ Thus, $\omega^2$ is an integrable $1$-form, and, because it is semi-basic for the submersion $x:B_1(S)\to S\subset\mathbb{R}^4$, it follows that $\omega^2$ is a multiple of the $x$-pullback of a (nonvanishing) $1$-form on $S$. Thus, $S$ is foliated by (connected) curves whose $x$-preimages in $B_1(S)$ are codimension $1$ integral submanifolds of $\omega^2$.</p> <p>I claim that these curves in $S$ are, in fact, lines in $\mathbb{R}^4$. To see this, note that, when one pulls back to a leaf of $\omega^2=0$ in $B_1(S)$, one has $\theta^2_1 = \theta^3_1=\theta^4_1= 0$ as well, so one has $$ \mathrm{d}x = e_1\,\omega^1\qquad\text{and}\qquad \mathrm{d}e_1 = e_1\,\theta^1_1\,. $$ In particular, the direction of $e_1$ is fixed on this leaf, and this is the tangent direction of the mapping $x$ restricted to this leaf. Hence the $x$-image of this leaf is an open interval in a line in $\mathbb{R}^4$. Thus, the surface is ruled, as claimed.<p></p> <p>There still remains the question of how one could 'generate' these surfaces, at least locally. I claim that, since these surfaces are, in fact, $1$-parameter families of lines, one should think of them as curves in the space of lines that satisfy some differential equations. Here is how one can think of this system of equations:<p></p> <p>Let $\mathcal{I}$ denote the Pfaffian system of rank $5$ that is generated by the five linearly independent $1$-forms $$ \omega^3,\ \omega^4,\ \theta^4_1\,,\ \theta^4_2\,,\ \theta^3_1 $$ (these are the $1$-forms that vanish when pulled back to $B_1(S)$ when $S\subset\mathbb{R}^4$ is a surface satisfying our hypotheses). By the structure equations, $$ \left. \begin{aligned} \mathrm{d}\omega^4 &amp;\equiv 0\\ \mathrm{d}\theta^4_1 &amp;\equiv 0\\ \mathrm{d}\omega^3 &amp;\equiv -\theta^3_2\wedge\omega^2\\ \mathrm{d}\theta^3_1 &amp;\equiv -\theta^3_2\wedge\theta^2_1\\ \mathrm{d}\theta^4_2 &amp;\equiv \phantom{-}\theta^3_2\wedge\theta^4_3\\ \end{aligned}\ \right\} \mathrm{modulo}\ \mathcal{I} $$ It follows that $\mathcal{J}$, the Cartan system of $\mathcal{I}$, has rank $9$ and is spanned by the nine $1$-forms $$ \omega^3,\ \omega^4,\ \theta^4_1\,,\ \theta^4_2\,,\ \theta^3_1\,,\ \omega^2,\ \theta^2_1\,,\ \theta^3_2\,,\ \theta^4_3 $$ In fact, $\mathcal{J}$ is easily seen to be the bundle of $1$-forms on $\mathbb{R}^4\times\mathrm{GL}(4,\mathbb{R})$ that is semibasic for the projection to the $9$-manifold $F$, where the projection sends $(p; v_i)$ to the element of $F$ described by $$ \bigl (\lambda(p,v_1),\ [v_1\wedge v_2], [v_1\wedge v_2\wedge v_3]\ \bigr) $$ and where $\lambda(p,v_1)$ is the line through $p$ in the direction $v_1$.</p> <p>In particular $\mathcal{I}$ is the pullback of a well-defined Pfaffian system of rank $5$ on $F$, whose annihilator is the $4$-plane field $D\subset TF$. Moreover, it is easy to show that the three rank $6$ Pfaffian systems generated by adjoining any one of $\omega^2$, $\theta^3_2$, or $\theta^4_3$ to $\mathcal{I}$ are themselves pullbacks of rank $6$ Pfaffian systems on $F$ whose annihilators in $TF$ are each $3$-plane subbundles of $D$. <p> By its very construction, the projection of $B_1(S)$ into $F$ is a curve that is tangent to $D$ and not tangent to any of these three $3$-plane subbundles of $D$.</p> <p>Conversely, if $\gamma\subset F$ is any integral curve of $D$ that is not tangent to any of the three $3$-plane subbundles of $D$, its preimage in $\mathbb{R}^4\times\mathrm{GL}(4,\mathbb{R})$ is a submanifold of the form $B_1(S)$ for a surface $S$ satisfying our conditions, in fact, the surface swept out by the union of the lines represented by $\lambda(\gamma)$, where $\lambda:F\to \Lambda$ is the obvious map to the lines.</p> <p>It is a standard fact that the curves tangent to a 4-plane field in a smooth manifold are locally described (up to reparametrization) by prescribing $3$ functions of one variable.</p>
How do i solve this : $\displaystyle \ f&#39;=e^{{f}^{-1}}$?
258,525
16
https://mathoverflow.net/questions/258525/how-do-i-solve-this-displaystyle-f-ef-1
<p>Let $f$ be a function such that :$f:\mathbb{R}\to \mathbb{R}$ and $f^{-1}$ is a compositional inverse of $f$. I would'd like to know how do I solve this class of differential equation : $$\displaystyle \ f'= e^{\displaystyle {f}^{-1}}?$$ </p> <p><strong>Note 01:</strong> $f' =\displaystyle\frac{df}{dx}$.</p> <p><strong>Edit:</strong> ${f}^{-1}$ is the inverse compositional of $f$, for example $\log$ is the inverse application of exp function .</p> <p><strong>Note 02</strong>: I have edited my question to clarify the titled question that related to ${f}^{-1}$</p> <p>Thank you for any help </p>
[ { "answer_id": 258544, "score": 36, "text": "<p>There is no such function. Since $f$ would have to map $\\mathbb R$ <em>onto</em> $\\mathbb R$ for the equation to make sense at all $x\\in\\mathbb R$, it follows that $f^{-1}(x)\\to -\\infty$ also as $x\\to -\\infty$, so $f'\\to 0$. Thus $f(x)\\ge x$, say, for all small enough $x$, hence $f^{-1}(x)\\le x$ eventually, but then the equation shows that $f'\\le e^x$, which is integrable on $(-\\infty, 0)$, so $f$ would approach a limit as $x\\to -\\infty$ and not be surjective after all.</p>\n" }, { "answer_id": 258695, "score": 5, "text": "<p>A formal Taylor series (e.g.f.) solution about the origin can be obtained a few ways.</p>\n\n<p>Let $f^{(-1)}(x) = e^{b.x}$ with $(b.)^n=b_n \\;$ and $ \\; b_0=0$.</p>\n\n<p>Then <a href=\"http://oeis.org/A036040\" rel=\"noreferrer\">A036040</a> (Bell polynomials) gives the e.g.f. </p>\n\n<p>$$e^{f^{(-1)}(x)}= e^{e^{b.x}}= 1 + b_1 x + (b_2+b_1^2) \\frac{x^2}{2!}+(b_3+3b_1b_2+b_1^3)\\frac{x^3}{3!}+\\cdots \\; ,$$</p>\n\n<p>and the Lagrange inversion / series reversion formula (LIF) <a href=\"http://oeis.org/A134685\" rel=\"noreferrer\">A134685</a> gives</p>\n\n<p>$$f'(x)= \\frac{1}{b_1} + \\frac{1}{b_1^3} (-b_2) x + \\frac{1}{b_1^5}(3b_2^2-b_1b_3)\\frac{x^2}{2!}+\\cdots \\; .$$</p>\n\n<p>Equating the two series and solving recursively gives</p>\n\n<p>$$b_n \\rightarrow (0,1,-1,3,-16,126,-1333,...)$$</p>\n\n<p>which is signed <a href=\"http://oeis.org/A214645\" rel=\"noreferrer\">A214645</a>. This follows from the application of the inverse function theorem (essentially the LIF again)</p>\n\n<p>$$f'(z) = 1/f^{(-1)}{'}(\\omega) \\; ,$$</p>\n\n<p>when $(z,\\omega)=(f^{(-1)}(\\omega),f(z)) $, leading to</p>\n\n<p>$$f^{(-1)}{'}(x) = \\exp[-f^{(-1)}(f^{(-1)}(x))],$$</p>\n\n<p>the differential equation defining signed A214645.</p>\n\n<p>Applying the LIF to the sequence for $b_n$ gives the e.g.f. $f(x)=e^{a.x}$ equivalent of F.C.'s o.g.f.</p>\n\n<p>$$ a_n \\rightarrow (0,1,1,0,1,-6,52,...).$$</p>\n\n<p>As another consistency check, apply the formalism of <a href=\"http://oeis.org/A133314\" rel=\"noreferrer\">A133314</a> for finding the multiplicative inverse of an e.g.f. to find the e.g.f. for $\\exp[-A(-x)]=\\exp[f^{(-1)}(x)]$ from that for </p>\n\n<p>$$\\exp[A(-x)]= 1 - x + 2 \\frac{x^2}{2!}-7 \\frac{x^3}{3!}+\\cdots \\; ,$$</p>\n\n<p>which is signed <a href=\"http://oeis.org/A233335\" rel=\"noreferrer\">A233335</a>, as noted in A214645. This gives $f'(x)=a. \\; e^{a.x}$.</p>\n" }, { "answer_id": 258544, "score": 36, "text": "<p>There is no such function. Since $f$ would have to map $\\mathbb R$ <em>onto</em> $\\mathbb R$ for the equation to make sense at all $x\\in\\mathbb R$, it follows that $f^{-1}(x)\\to -\\infty$ also as $x\\to -\\infty$, so $f'\\to 0$. Thus $f(x)\\ge x$, say, for all small enough $x$, hence $f^{-1}(x)\\le x$ eventually, but then the equation shows that $f'\\le e^x$, which is integrable on $(-\\infty, 0)$, so $f$ would approach a limit as $x\\to -\\infty$ and not be surjective after all.</p>\n" }, { "answer_id": 258695, "score": 5, "text": "<p>A formal Taylor series (e.g.f.) solution about the origin can be obtained a few ways.</p>\n\n<p>Let $f^{(-1)}(x) = e^{b.x}$ with $(b.)^n=b_n \\;$ and $ \\; b_0=0$.</p>\n\n<p>Then <a href=\"http://oeis.org/A036040\" rel=\"noreferrer\">A036040</a> (Bell polynomials) gives the e.g.f. </p>\n\n<p>$$e^{f^{(-1)}(x)}= e^{e^{b.x}}= 1 + b_1 x + (b_2+b_1^2) \\frac{x^2}{2!}+(b_3+3b_1b_2+b_1^3)\\frac{x^3}{3!}+\\cdots \\; ,$$</p>\n\n<p>and the Lagrange inversion / series reversion formula (LIF) <a href=\"http://oeis.org/A134685\" rel=\"noreferrer\">A134685</a> gives</p>\n\n<p>$$f'(x)= \\frac{1}{b_1} + \\frac{1}{b_1^3} (-b_2) x + \\frac{1}{b_1^5}(3b_2^2-b_1b_3)\\frac{x^2}{2!}+\\cdots \\; .$$</p>\n\n<p>Equating the two series and solving recursively gives</p>\n\n<p>$$b_n \\rightarrow (0,1,-1,3,-16,126,-1333,...)$$</p>\n\n<p>which is signed <a href=\"http://oeis.org/A214645\" rel=\"noreferrer\">A214645</a>. This follows from the application of the inverse function theorem (essentially the LIF again)</p>\n\n<p>$$f'(z) = 1/f^{(-1)}{'}(\\omega) \\; ,$$</p>\n\n<p>when $(z,\\omega)=(f^{(-1)}(\\omega),f(z)) $, leading to</p>\n\n<p>$$f^{(-1)}{'}(x) = \\exp[-f^{(-1)}(f^{(-1)}(x))],$$</p>\n\n<p>the differential equation defining signed A214645.</p>\n\n<p>Applying the LIF to the sequence for $b_n$ gives the e.g.f. $f(x)=e^{a.x}$ equivalent of F.C.'s o.g.f.</p>\n\n<p>$$ a_n \\rightarrow (0,1,1,0,1,-6,52,...).$$</p>\n\n<p>As another consistency check, apply the formalism of <a href=\"http://oeis.org/A133314\" rel=\"noreferrer\">A133314</a> for finding the multiplicative inverse of an e.g.f. to find the e.g.f. for $\\exp[-A(-x)]=\\exp[f^{(-1)}(x)]$ from that for </p>\n\n<p>$$\\exp[A(-x)]= 1 - x + 2 \\frac{x^2}{2!}-7 \\frac{x^3}{3!}+\\cdots \\; ,$$</p>\n\n<p>which is signed <a href=\"http://oeis.org/A233335\" rel=\"noreferrer\">A233335</a>, as noted in A214645. This gives $f'(x)=a. \\; e^{a.x}$.</p>\n" }, { "answer_id": 258544, "score": 36, "text": "<p>There is no such function. Since $f$ would have to map $\\mathbb R$ <em>onto</em> $\\mathbb R$ for the equation to make sense at all $x\\in\\mathbb R$, it follows that $f^{-1}(x)\\to -\\infty$ also as $x\\to -\\infty$, so $f'\\to 0$. Thus $f(x)\\ge x$, say, for all small enough $x$, hence $f^{-1}(x)\\le x$ eventually, but then the equation shows that $f'\\le e^x$, which is integrable on $(-\\infty, 0)$, so $f$ would approach a limit as $x\\to -\\infty$ and not be surjective after all.</p>\n" }, { "answer_id": 258695, "score": 5, "text": "<p>A formal Taylor series (e.g.f.) solution about the origin can be obtained a few ways.</p>\n\n<p>Let $f^{(-1)}(x) = e^{b.x}$ with $(b.)^n=b_n \\;$ and $ \\; b_0=0$.</p>\n\n<p>Then <a href=\"http://oeis.org/A036040\" rel=\"noreferrer\">A036040</a> (Bell polynomials) gives the e.g.f. </p>\n\n<p>$$e^{f^{(-1)}(x)}= e^{e^{b.x}}= 1 + b_1 x + (b_2+b_1^2) \\frac{x^2}{2!}+(b_3+3b_1b_2+b_1^3)\\frac{x^3}{3!}+\\cdots \\; ,$$</p>\n\n<p>and the Lagrange inversion / series reversion formula (LIF) <a href=\"http://oeis.org/A134685\" rel=\"noreferrer\">A134685</a> gives</p>\n\n<p>$$f'(x)= \\frac{1}{b_1} + \\frac{1}{b_1^3} (-b_2) x + \\frac{1}{b_1^5}(3b_2^2-b_1b_3)\\frac{x^2}{2!}+\\cdots \\; .$$</p>\n\n<p>Equating the two series and solving recursively gives</p>\n\n<p>$$b_n \\rightarrow (0,1,-1,3,-16,126,-1333,...)$$</p>\n\n<p>which is signed <a href=\"http://oeis.org/A214645\" rel=\"noreferrer\">A214645</a>. This follows from the application of the inverse function theorem (essentially the LIF again)</p>\n\n<p>$$f'(z) = 1/f^{(-1)}{'}(\\omega) \\; ,$$</p>\n\n<p>when $(z,\\omega)=(f^{(-1)}(\\omega),f(z)) $, leading to</p>\n\n<p>$$f^{(-1)}{'}(x) = \\exp[-f^{(-1)}(f^{(-1)}(x))],$$</p>\n\n<p>the differential equation defining signed A214645.</p>\n\n<p>Applying the LIF to the sequence for $b_n$ gives the e.g.f. $f(x)=e^{a.x}$ equivalent of F.C.'s o.g.f.</p>\n\n<p>$$ a_n \\rightarrow (0,1,1,0,1,-6,52,...).$$</p>\n\n<p>As another consistency check, apply the formalism of <a href=\"http://oeis.org/A133314\" rel=\"noreferrer\">A133314</a> for finding the multiplicative inverse of an e.g.f. to find the e.g.f. for $\\exp[-A(-x)]=\\exp[f^{(-1)}(x)]$ from that for </p>\n\n<p>$$\\exp[A(-x)]= 1 - x + 2 \\frac{x^2}{2!}-7 \\frac{x^3}{3!}+\\cdots \\; ,$$</p>\n\n<p>which is signed <a href=\"http://oeis.org/A233335\" rel=\"noreferrer\">A233335</a>, as noted in A214645. This gives $f'(x)=a. \\; e^{a.x}$.</p>\n" } ]
Question: How do i solve this : $\displaystyle \ f&#39;=e^{{f}^{-1}}$? Answer 1: <p>There is no such function. Since $f$ would have to map $\mathbb R$ <em>onto</em> $\mathbb R$ for the equation to make sense at all $x\in\mathbb R$, it follows that $f^{-1}(x)\to -\infty$ also as $x\to -\infty$, so $f'\to 0$. Thus $f(x)\ge x$, say, for all small enough $x$, hence $f^{-1}(x)\le x$ eventually, but then the equation shows that $f'\le e^x$, which is integrable on $(-\infty, 0)$, so $f$ would approach a limit as $x\to -\infty$ and not be surjective after all.</p> Answer 2: <p>A formal Taylor series (e.g.f.) solution about the origin can be obtained a few ways.</p> <p>Let $f^{(-1)}(x) = e^{b.x}$ with $(b.)^n=b_n \;$ and $ \; b_0=0$.</p> <p>Then <a href="http://oeis.org/A036040" rel="noreferrer">A036040</a> (Bell polynomials) gives the e.g.f. </p> <p>$$e^{f^{(-1)}(x)}= e^{e^{b.x}}= 1 + b_1 x + (b_2+b_1^2) \frac{x^2}{2!}+(b_3+3b_1b_2+b_1^3)\frac{x^3}{3!}+\cdots \; ,$$</p> <p>and the Lagrange inversion / series reversion formula (LIF) <a href="http://oeis.org/A134685" rel="noreferrer">A134685</a> gives</p> <p>$$f'(x)= \frac{1}{b_1} + \frac{1}{b_1^3} (-b_2) x + \frac{1}{b_1^5}(3b_2^2-b_1b_3)\frac{x^2}{2!}+\cdots \; .$$</p> <p>Equating the two series and solving recursively gives</p> <p>$$b_n \rightarrow (0,1,-1,3,-16,126,-1333,...)$$</p> <p>which is signed <a href="http://oeis.org/A214645" rel="noreferrer">A214645</a>. This follows from the application of the inverse function theorem (essentially the LIF again)</p> <p>$$f'(z) = 1/f^{(-1)}{'}(\omega) \; ,$$</p> <p>when $(z,\omega)=(f^{(-1)}(\omega),f(z)) $, leading to</p> <p>$$f^{(-1)}{'}(x) = \exp[-f^{(-1)}(f^{(-1)}(x))],$$</p> <p>the differential equation defining signed A214645.</p> <p>Applying the LIF to the sequence for $b_n$ gives the e.g.f. $f(x)=e^{a.x}$ equivalent of F.C.'s o.g.f.</p> <p>$$ a_n \rightarrow (0,1,1,0,1,-6,52,...).$$</p> <p>As another consistency check, apply the formalism of <a href="http://oeis.org/A133314" rel="noreferrer">A133314</a> for finding the multiplicative inverse of an e.g.f. to find the e.g.f. for $\exp[-A(-x)]=\exp[f^{(-1)}(x)]$ from that for </p> <p>$$\exp[A(-x)]= 1 - x + 2 \frac{x^2}{2!}-7 \frac{x^3}{3!}+\cdots \; ,$$</p> <p>which is signed <a href="http://oeis.org/A233335" rel="noreferrer">A233335</a>, as noted in A214645. This gives $f'(x)=a. \; e^{a.x}$.</p> Answer 3: <p>There is no such function. Since $f$ would have to map $\mathbb R$ <em>onto</em> $\mathbb R$ for the equation to make sense at all $x\in\mathbb R$, it follows that $f^{-1}(x)\to -\infty$ also as $x\to -\infty$, so $f'\to 0$. Thus $f(x)\ge x$, say, for all small enough $x$, hence $f^{-1}(x)\le x$ eventually, but then the equation shows that $f'\le e^x$, which is integrable on $(-\infty, 0)$, so $f$ would approach a limit as $x\to -\infty$ and not be surjective after all.</p> Answer 4: <p>A formal Taylor series (e.g.f.) solution about the origin can be obtained a few ways.</p> <p>Let $f^{(-1)}(x) = e^{b.x}$ with $(b.)^n=b_n \;$ and $ \; b_0=0$.</p> <p>Then <a href="http://oeis.org/A036040" rel="noreferrer">A036040</a> (Bell polynomials) gives the e.g.f. </p> <p>$$e^{f^{(-1)}(x)}= e^{e^{b.x}}= 1 + b_1 x + (b_2+b_1^2) \frac{x^2}{2!}+(b_3+3b_1b_2+b_1^3)\frac{x^3}{3!}+\cdots \; ,$$</p> <p>and the Lagrange inversion / series reversion formula (LIF) <a href="http://oeis.org/A134685" rel="noreferrer">A134685</a> gives</p> <p>$$f'(x)= \frac{1}{b_1} + \frac{1}{b_1^3} (-b_2) x + \frac{1}{b_1^5}(3b_2^2-b_1b_3)\frac{x^2}{2!}+\cdots \; .$$</p> <p>Equating the two series and solving recursively gives</p> <p>$$b_n \rightarrow (0,1,-1,3,-16,126,-1333,...)$$</p> <p>which is signed <a href="http://oeis.org/A214645" rel="noreferrer">A214645</a>. This follows from the application of the inverse function theorem (essentially the LIF again)</p> <p>$$f'(z) = 1/f^{(-1)}{'}(\omega) \; ,$$</p> <p>when $(z,\omega)=(f^{(-1)}(\omega),f(z)) $, leading to</p> <p>$$f^{(-1)}{'}(x) = \exp[-f^{(-1)}(f^{(-1)}(x))],$$</p> <p>the differential equation defining signed A214645.</p> <p>Applying the LIF to the sequence for $b_n$ gives the e.g.f. $f(x)=e^{a.x}$ equivalent of F.C.'s o.g.f.</p> <p>$$ a_n \rightarrow (0,1,1,0,1,-6,52,...).$$</p> <p>As another consistency check, apply the formalism of <a href="http://oeis.org/A133314" rel="noreferrer">A133314</a> for finding the multiplicative inverse of an e.g.f. to find the e.g.f. for $\exp[-A(-x)]=\exp[f^{(-1)}(x)]$ from that for </p> <p>$$\exp[A(-x)]= 1 - x + 2 \frac{x^2}{2!}-7 \frac{x^3}{3!}+\cdots \; ,$$</p> <p>which is signed <a href="http://oeis.org/A233335" rel="noreferrer">A233335</a>, as noted in A214645. This gives $f'(x)=a. \; e^{a.x}$.</p> Answer 5: <p>There is no such function. Since $f$ would have to map $\mathbb R$ <em>onto</em> $\mathbb R$ for the equation to make sense at all $x\in\mathbb R$, it follows that $f^{-1}(x)\to -\infty$ also as $x\to -\infty$, so $f'\to 0$. Thus $f(x)\ge x$, say, for all small enough $x$, hence $f^{-1}(x)\le x$ eventually, but then the equation shows that $f'\le e^x$, which is integrable on $(-\infty, 0)$, so $f$ would approach a limit as $x\to -\infty$ and not be surjective after all.</p> Answer 6: <p>A formal Taylor series (e.g.f.) solution about the origin can be obtained a few ways.</p> <p>Let $f^{(-1)}(x) = e^{b.x}$ with $(b.)^n=b_n \;$ and $ \; b_0=0$.</p> <p>Then <a href="http://oeis.org/A036040" rel="noreferrer">A036040</a> (Bell polynomials) gives the e.g.f. </p> <p>$$e^{f^{(-1)}(x)}= e^{e^{b.x}}= 1 + b_1 x + (b_2+b_1^2) \frac{x^2}{2!}+(b_3+3b_1b_2+b_1^3)\frac{x^3}{3!}+\cdots \; ,$$</p> <p>and the Lagrange inversion / series reversion formula (LIF) <a href="http://oeis.org/A134685" rel="noreferrer">A134685</a> gives</p> <p>$$f'(x)= \frac{1}{b_1} + \frac{1}{b_1^3} (-b_2) x + \frac{1}{b_1^5}(3b_2^2-b_1b_3)\frac{x^2}{2!}+\cdots \; .$$</p> <p>Equating the two series and solving recursively gives</p> <p>$$b_n \rightarrow (0,1,-1,3,-16,126,-1333,...)$$</p> <p>which is signed <a href="http://oeis.org/A214645" rel="noreferrer">A214645</a>. This follows from the application of the inverse function theorem (essentially the LIF again)</p> <p>$$f'(z) = 1/f^{(-1)}{'}(\omega) \; ,$$</p> <p>when $(z,\omega)=(f^{(-1)}(\omega),f(z)) $, leading to</p> <p>$$f^{(-1)}{'}(x) = \exp[-f^{(-1)}(f^{(-1)}(x))],$$</p> <p>the differential equation defining signed A214645.</p> <p>Applying the LIF to the sequence for $b_n$ gives the e.g.f. $f(x)=e^{a.x}$ equivalent of F.C.'s o.g.f.</p> <p>$$ a_n \rightarrow (0,1,1,0,1,-6,52,...).$$</p> <p>As another consistency check, apply the formalism of <a href="http://oeis.org/A133314" rel="noreferrer">A133314</a> for finding the multiplicative inverse of an e.g.f. to find the e.g.f. for $\exp[-A(-x)]=\exp[f^{(-1)}(x)]$ from that for </p> <p>$$\exp[A(-x)]= 1 - x + 2 \frac{x^2}{2!}-7 \frac{x^3}{3!}+\cdots \; ,$$</p> <p>which is signed <a href="http://oeis.org/A233335" rel="noreferrer">A233335</a>, as noted in A214645. This gives $f'(x)=a. \; e^{a.x}$.</p>
Can a convex polytope with $f$ facets have more than $f$ facets when projected into $\mathbb{R}^2$?
260,107
15
https://mathoverflow.net/questions/260107/can-a-convex-polytope-with-f-facets-have-more-than-f-facets-when-projected-i
<p>Let $P$ be a convex polytope in $\mathbb{R}^d$ with $n$ vertices and $f$ facets. Let $\text{Proj}(P)$ denote the projection of $P$ into $\mathbb{R}^2$.</p> <p>Can $\text{Proj}(P)$ have more than $f$ facets?</p> <p>In the general case, each successive projection can increase the number of facets from $f$ to $\left\lfloor \frac{f^2}{2} \right\rfloor$, but I'm wondering if $\mathbb{R}^2$ is a special case.</p>
[ { "answer_id": 260114, "score": 25, "text": "<p>Consider the polytope in $\\mathbb{R}^3$ with $8$ vertices at coordinates $(\\pm 1, \\pm 2, 1), (\\pm 2, \\pm 1, -1)$. Geometrically this looks like a cube where the top face is stretched in the direction of the $y$-axis and the bottom face is stretched in the direction of the $x$-axis, but it still has the face structure of a cube and in particular has $6$ facets.</p>\n\n<p>It's projection onto the first two coordinates is clearly an octagon, and hence has two more facets than the original polytope.</p>\n" }, { "answer_id": 260175, "score": 25, "text": "<p>Your question is essentially about extension complexity. In general, the <em>extension complexity</em> of a polytope <span class=\"math-container\">$P$</span> is the minimum number of facets over all polytopes <span class=\"math-container\">$Q$</span> which project to <span class=\"math-container\">$P$</span>. You are interested in the extension complexity of polygons. <a href=\"https://arxiv.org/abs/1107.0371\" rel=\"noreferrer\">Fiorini, Rothvoß, and Tiwary</a> proved that <em>regular</em> <span class=\"math-container\">$n$</span>-gons have extension complexity <span class=\"math-container\">$O(\\log n)$</span>. For lower bounds, they give examples of <span class=\"math-container\">$n$</span>-gons which have extension complexity <span class=\"math-container\">$\\sqrt{2n}$</span>.</p>\n<p>It is an <a href=\"http://www.openproblemgarden.org/op/extension_complexity_of_convex_polygons\" rel=\"noreferrer\">open question</a> if there exists for infinitely many <span class=\"math-container\">$n$</span>, an <span class=\"math-container\">$n$</span>-gon with extension complexity <span class=\"math-container\">$\\Omega(n)$</span>.\n<strong>Edit.</strong> <a href=\"https://arxiv.org/abs/1412.0728\" rel=\"noreferrer\">Shitov</a> has recently shown that every convex <span class=\"math-container\">$n$</span>-gon has extension complexity at most <span class=\"math-container\">$147 n^{2/3}$</span>, which resolves the open question in the negative.</p>\n<p>As a bonus, here is a picture of the polytope in Nate's answer (courtesy of Samuel Fiorini).</p>\n<p><a href=\"https://i.stack.imgur.com/5RiRT.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/5RiRT.png\" alt=\"enter image description here\" /></a></p>\n" }, { "answer_id": 260157, "score": 16, "text": "<p>Another, combinatorially minimal, counterexample of such a polytope $P$ (with only five facets) is the convex hull of the six vertices $(\\pm2, 0, 0)$, $(\\pm1, \\pm1, 1)$. Its projection to the $xy$-plane is a hexagon. The minimality of the number of vertices and the number of facets is easy to prove.</p>\n" }, { "answer_id": 260114, "score": 25, "text": "<p>Consider the polytope in $\\mathbb{R}^3$ with $8$ vertices at coordinates $(\\pm 1, \\pm 2, 1), (\\pm 2, \\pm 1, -1)$. Geometrically this looks like a cube where the top face is stretched in the direction of the $y$-axis and the bottom face is stretched in the direction of the $x$-axis, but it still has the face structure of a cube and in particular has $6$ facets.</p>\n\n<p>It's projection onto the first two coordinates is clearly an octagon, and hence has two more facets than the original polytope.</p>\n" }, { "answer_id": 260175, "score": 25, "text": "<p>Your question is essentially about extension complexity. In general, the <em>extension complexity</em> of a polytope <span class=\"math-container\">$P$</span> is the minimum number of facets over all polytopes <span class=\"math-container\">$Q$</span> which project to <span class=\"math-container\">$P$</span>. You are interested in the extension complexity of polygons. <a href=\"https://arxiv.org/abs/1107.0371\" rel=\"noreferrer\">Fiorini, Rothvoß, and Tiwary</a> proved that <em>regular</em> <span class=\"math-container\">$n$</span>-gons have extension complexity <span class=\"math-container\">$O(\\log n)$</span>. For lower bounds, they give examples of <span class=\"math-container\">$n$</span>-gons which have extension complexity <span class=\"math-container\">$\\sqrt{2n}$</span>.</p>\n<p>It is an <a href=\"http://www.openproblemgarden.org/op/extension_complexity_of_convex_polygons\" rel=\"noreferrer\">open question</a> if there exists for infinitely many <span class=\"math-container\">$n$</span>, an <span class=\"math-container\">$n$</span>-gon with extension complexity <span class=\"math-container\">$\\Omega(n)$</span>.\n<strong>Edit.</strong> <a href=\"https://arxiv.org/abs/1412.0728\" rel=\"noreferrer\">Shitov</a> has recently shown that every convex <span class=\"math-container\">$n$</span>-gon has extension complexity at most <span class=\"math-container\">$147 n^{2/3}$</span>, which resolves the open question in the negative.</p>\n<p>As a bonus, here is a picture of the polytope in Nate's answer (courtesy of Samuel Fiorini).</p>\n<p><a href=\"https://i.stack.imgur.com/5RiRT.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/5RiRT.png\" alt=\"enter image description here\" /></a></p>\n" }, { "answer_id": 260157, "score": 16, "text": "<p>Another, combinatorially minimal, counterexample of such a polytope $P$ (with only five facets) is the convex hull of the six vertices $(\\pm2, 0, 0)$, $(\\pm1, \\pm1, 1)$. Its projection to the $xy$-plane is a hexagon. The minimality of the number of vertices and the number of facets is easy to prove.</p>\n" }, { "answer_id": 260114, "score": 25, "text": "<p>Consider the polytope in $\\mathbb{R}^3$ with $8$ vertices at coordinates $(\\pm 1, \\pm 2, 1), (\\pm 2, \\pm 1, -1)$. Geometrically this looks like a cube where the top face is stretched in the direction of the $y$-axis and the bottom face is stretched in the direction of the $x$-axis, but it still has the face structure of a cube and in particular has $6$ facets.</p>\n\n<p>It's projection onto the first two coordinates is clearly an octagon, and hence has two more facets than the original polytope.</p>\n" }, { "answer_id": 260175, "score": 25, "text": "<p>Your question is essentially about extension complexity. In general, the <em>extension complexity</em> of a polytope <span class=\"math-container\">$P$</span> is the minimum number of facets over all polytopes <span class=\"math-container\">$Q$</span> which project to <span class=\"math-container\">$P$</span>. You are interested in the extension complexity of polygons. <a href=\"https://arxiv.org/abs/1107.0371\" rel=\"noreferrer\">Fiorini, Rothvoß, and Tiwary</a> proved that <em>regular</em> <span class=\"math-container\">$n$</span>-gons have extension complexity <span class=\"math-container\">$O(\\log n)$</span>. For lower bounds, they give examples of <span class=\"math-container\">$n$</span>-gons which have extension complexity <span class=\"math-container\">$\\sqrt{2n}$</span>.</p>\n<p>It is an <a href=\"http://www.openproblemgarden.org/op/extension_complexity_of_convex_polygons\" rel=\"noreferrer\">open question</a> if there exists for infinitely many <span class=\"math-container\">$n$</span>, an <span class=\"math-container\">$n$</span>-gon with extension complexity <span class=\"math-container\">$\\Omega(n)$</span>.\n<strong>Edit.</strong> <a href=\"https://arxiv.org/abs/1412.0728\" rel=\"noreferrer\">Shitov</a> has recently shown that every convex <span class=\"math-container\">$n$</span>-gon has extension complexity at most <span class=\"math-container\">$147 n^{2/3}$</span>, which resolves the open question in the negative.</p>\n<p>As a bonus, here is a picture of the polytope in Nate's answer (courtesy of Samuel Fiorini).</p>\n<p><a href=\"https://i.stack.imgur.com/5RiRT.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/5RiRT.png\" alt=\"enter image description here\" /></a></p>\n" }, { "answer_id": 260157, "score": 16, "text": "<p>Another, combinatorially minimal, counterexample of such a polytope $P$ (with only five facets) is the convex hull of the six vertices $(\\pm2, 0, 0)$, $(\\pm1, \\pm1, 1)$. Its projection to the $xy$-plane is a hexagon. The minimality of the number of vertices and the number of facets is easy to prove.</p>\n" } ]
Question: Can a convex polytope with $f$ facets have more than $f$ facets when projected into $\mathbb{R}^2$? Answer 1: <p>Consider the polytope in $\mathbb{R}^3$ with $8$ vertices at coordinates $(\pm 1, \pm 2, 1), (\pm 2, \pm 1, -1)$. Geometrically this looks like a cube where the top face is stretched in the direction of the $y$-axis and the bottom face is stretched in the direction of the $x$-axis, but it still has the face structure of a cube and in particular has $6$ facets.</p> <p>It's projection onto the first two coordinates is clearly an octagon, and hence has two more facets than the original polytope.</p> Answer 2: <p>Your question is essentially about extension complexity. In general, the <em>extension complexity</em> of a polytope <span class="math-container">$P$</span> is the minimum number of facets over all polytopes <span class="math-container">$Q$</span> which project to <span class="math-container">$P$</span>. You are interested in the extension complexity of polygons. <a href="https://arxiv.org/abs/1107.0371" rel="noreferrer">Fiorini, Rothvoß, and Tiwary</a> proved that <em>regular</em> <span class="math-container">$n$</span>-gons have extension complexity <span class="math-container">$O(\log n)$</span>. For lower bounds, they give examples of <span class="math-container">$n$</span>-gons which have extension complexity <span class="math-container">$\sqrt{2n}$</span>.</p> <p>It is an <a href="http://www.openproblemgarden.org/op/extension_complexity_of_convex_polygons" rel="noreferrer">open question</a> if there exists for infinitely many <span class="math-container">$n$</span>, an <span class="math-container">$n$</span>-gon with extension complexity <span class="math-container">$\Omega(n)$</span>. <strong>Edit.</strong> <a href="https://arxiv.org/abs/1412.0728" rel="noreferrer">Shitov</a> has recently shown that every convex <span class="math-container">$n$</span>-gon has extension complexity at most <span class="math-container">$147 n^{2/3}$</span>, which resolves the open question in the negative.</p> <p>As a bonus, here is a picture of the polytope in Nate's answer (courtesy of Samuel Fiorini).</p> <p><a href="https://i.stack.imgur.com/5RiRT.png" rel="noreferrer"><img src="https://i.stack.imgur.com/5RiRT.png" alt="enter image description here" /></a></p> Answer 3: <p>Another, combinatorially minimal, counterexample of such a polytope $P$ (with only five facets) is the convex hull of the six vertices $(\pm2, 0, 0)$, $(\pm1, \pm1, 1)$. Its projection to the $xy$-plane is a hexagon. The minimality of the number of vertices and the number of facets is easy to prove.</p> Answer 4: <p>Consider the polytope in $\mathbb{R}^3$ with $8$ vertices at coordinates $(\pm 1, \pm 2, 1), (\pm 2, \pm 1, -1)$. Geometrically this looks like a cube where the top face is stretched in the direction of the $y$-axis and the bottom face is stretched in the direction of the $x$-axis, but it still has the face structure of a cube and in particular has $6$ facets.</p> <p>It's projection onto the first two coordinates is clearly an octagon, and hence has two more facets than the original polytope.</p> Answer 5: <p>Your question is essentially about extension complexity. In general, the <em>extension complexity</em> of a polytope <span class="math-container">$P$</span> is the minimum number of facets over all polytopes <span class="math-container">$Q$</span> which project to <span class="math-container">$P$</span>. You are interested in the extension complexity of polygons. <a href="https://arxiv.org/abs/1107.0371" rel="noreferrer">Fiorini, Rothvoß, and Tiwary</a> proved that <em>regular</em> <span class="math-container">$n$</span>-gons have extension complexity <span class="math-container">$O(\log n)$</span>. For lower bounds, they give examples of <span class="math-container">$n$</span>-gons which have extension complexity <span class="math-container">$\sqrt{2n}$</span>.</p> <p>It is an <a href="http://www.openproblemgarden.org/op/extension_complexity_of_convex_polygons" rel="noreferrer">open question</a> if there exists for infinitely many <span class="math-container">$n$</span>, an <span class="math-container">$n$</span>-gon with extension complexity <span class="math-container">$\Omega(n)$</span>. <strong>Edit.</strong> <a href="https://arxiv.org/abs/1412.0728" rel="noreferrer">Shitov</a> has recently shown that every convex <span class="math-container">$n$</span>-gon has extension complexity at most <span class="math-container">$147 n^{2/3}$</span>, which resolves the open question in the negative.</p> <p>As a bonus, here is a picture of the polytope in Nate's answer (courtesy of Samuel Fiorini).</p> <p><a href="https://i.stack.imgur.com/5RiRT.png" rel="noreferrer"><img src="https://i.stack.imgur.com/5RiRT.png" alt="enter image description here" /></a></p> Answer 6: <p>Another, combinatorially minimal, counterexample of such a polytope $P$ (with only five facets) is the convex hull of the six vertices $(\pm2, 0, 0)$, $(\pm1, \pm1, 1)$. Its projection to the $xy$-plane is a hexagon. The minimality of the number of vertices and the number of facets is easy to prove.</p> Answer 7: <p>Consider the polytope in $\mathbb{R}^3$ with $8$ vertices at coordinates $(\pm 1, \pm 2, 1), (\pm 2, \pm 1, -1)$. Geometrically this looks like a cube where the top face is stretched in the direction of the $y$-axis and the bottom face is stretched in the direction of the $x$-axis, but it still has the face structure of a cube and in particular has $6$ facets.</p> <p>It's projection onto the first two coordinates is clearly an octagon, and hence has two more facets than the original polytope.</p> Answer 8: <p>Your question is essentially about extension complexity. In general, the <em>extension complexity</em> of a polytope <span class="math-container">$P$</span> is the minimum number of facets over all polytopes <span class="math-container">$Q$</span> which project to <span class="math-container">$P$</span>. You are interested in the extension complexity of polygons. <a href="https://arxiv.org/abs/1107.0371" rel="noreferrer">Fiorini, Rothvoß, and Tiwary</a> proved that <em>regular</em> <span class="math-container">$n$</span>-gons have extension complexity <span class="math-container">$O(\log n)$</span>. For lower bounds, they give examples of <span class="math-container">$n$</span>-gons which have extension complexity <span class="math-container">$\sqrt{2n}$</span>.</p> <p>It is an <a href="http://www.openproblemgarden.org/op/extension_complexity_of_convex_polygons" rel="noreferrer">open question</a> if there exists for infinitely many <span class="math-container">$n$</span>, an <span class="math-container">$n$</span>-gon with extension complexity <span class="math-container">$\Omega(n)$</span>. <strong>Edit.</strong> <a href="https://arxiv.org/abs/1412.0728" rel="noreferrer">Shitov</a> has recently shown that every convex <span class="math-container">$n$</span>-gon has extension complexity at most <span class="math-container">$147 n^{2/3}$</span>, which resolves the open question in the negative.</p> <p>As a bonus, here is a picture of the polytope in Nate's answer (courtesy of Samuel Fiorini).</p> <p><a href="https://i.stack.imgur.com/5RiRT.png" rel="noreferrer"><img src="https://i.stack.imgur.com/5RiRT.png" alt="enter image description here" /></a></p> Answer 9: <p>Another, combinatorially minimal, counterexample of such a polytope $P$ (with only five facets) is the convex hull of the six vertices $(\pm2, 0, 0)$, $(\pm1, \pm1, 1)$. Its projection to the $xy$-plane is a hexagon. The minimality of the number of vertices and the number of facets is easy to prove.</p>
Modular forms from counting points on algebraic varieties over a finite field
260,156
15
https://mathoverflow.net/questions/260156/modular-forms-from-counting-points-on-algebraic-varieties-over-a-finite-field
<p>Suppose we are given some polynomial with integer coefficients, which we regard as carving out an affine variety $E$, for example:</p> <p>$$ 3x^2y - 12 x^3y^5 + 27y^9 - 2 = 0 \tag{$*$} $$</p> <p>(We might consider a bunch of equations, we might work over projective space, but let's keep it simple for now).</p> <p>We are interested in the number of points on $E$ when we reduce modulo $p$, i.e. over the finite field $\mathbb{F}_p$ as the prime $p$ varies. For our single equation in two variables, as a rough approximation, we would expect $p$ points in general. So for each prime $p$, we define numbers $a_p$ which measure the deviation from this, </p> <p>$$a_p := p - \text{number of solutions to $(*)$ over $\mathbb{F}_p$}$$</p> <p>Question: When is it true that the numbers $\{a_p\}$ are "modular", in the sense that there exists a modular form </p> <p>$$f = \sum_{n=1}^\infty b_n q^n$$</p> <p>such that $b_p = a_p$ for almost all primes $p$? (The "almost all" is to avoid problems with bad primes. Note that $f$ is still uniquely determined by the above requirement.)</p> <p>The Modularity Theorem of Breuil, Conrad, Taylor and Diamond says that this is true when $E$ is an elliptic curve, i.e. takes the form $y^2 = 4x^3 - g_2 x - g_3$ for some integers $g_2, g_3$. In that case, $f$ is a weight 2 modular form of level $N$ where $N$ is the "conductor" of $E$. </p> <p>But is it true for more general varieties? </p> <p>(Note: I am aware of a generalized "Modularity Theorem" for certain Abelian varieties, but it's not clear to me that what people mean by "Modular" in that context is the same as the simple-minded notion I'm using --- that an adjusted count of points mod $p$ gives the Fourier coefficients of a modular form.) </p>
[ { "answer_id": 260232, "score": 22, "text": "<p>The correct setting for this construction turns out to be <em>projective</em> varieties, so let me suppose we have a smooth variety $X$ inside $\\mathbf{P}^N$, for some $N \\ge 1$, defined by the vanishing of some homogenous polynomials $F_1, \\dots, F_r$ in variables $x_0, \\dots, x_N$, with the $F_i$ having coefficients in $\\mathbf{Q}$. Actually, let me assume the $F_i$ have coefficients in $\\mathbf{Z}$, which is no loss since we can just multiply up. Then we can make sense of the reduction of $X$ modulo $p$; and we want to study the point counts $\\#X(\\mathbf{F}_p)$ as a function of $p$, possibly neglecting some finite set $\\Sigma$ containing all primes such that the reduction of $X$ mod $p$ is singular.</p>\n\n<p>Thanks to Grothendieck, Deligne, and others, we have a very powerful bunch of tools for analysing this problem. The setup is as follows. Choose your favourite prime $\\ell$. Then the theory of <strong>etale cohomology</strong> attaches to $X$ a bunch of finite-dimensional $\\mathbf{Q}_\\ell$-vector spaces $H^i_{\\mathrm{et}}(X_{\\overline{\\mathbf{Q}}}, \\mathbf{Q}_\\ell)$ (let me abbreviate this by $H^i_\\ell(X)$ to save typing). The dimension of $H^i_\\ell$ is the same as the $i$-th Betti number of the manifold $X(\\mathbb{C})$; but they encode much more data, because each $H^i_\\ell(X)$ is a representation of the Galois group $\\operatorname{Gal}(\\overline{\\mathbf{Q}} / \\mathbf{Q})$, unramified outside $\\Sigma \\cup \\{\\ell\\}$; so for every prime not in this set, and every $i$, we have a number $t_i(p)$, the trace of Frobenius at $p$ on $H^i_\\ell$, which turns out to be independent of $\\ell$.</p>\n\n<p><strong>Theorem</strong>: $\\#X(\\mathbf{F}_p)$ is the alternating sum $\\sum_{i=0}^{2 dim(X)} (-1)^i t_i(p)$.</p>\n\n<p>Now let's analyse $H^i_\\ell$ as a Galois representation. Representations of Galois groups needn't be direct sums of irreducibles, but we can replace $H^i_\\ell$ by its semisimplification, which does have this property and has the same trace as the original $H^i_\\ell$. This semisimplification will look like $V_{i, 1} + \\dots + V_{i, r_i}$ where $V_{i, j}$ are irreducible; and the $V_{i, j}$ all have motivic weight $i$, so the same $V$ can't appear for two different $i$'s. So we get a slightly finer decomposition</p>\n\n<p>$\\#X(\\mathbf{F}_p) = \\sum_{i=0}^{2 \\mathrm{dim} X} (-1)^i \\sum_{j=1}^{k_i} t_{i, j}(p)$</p>\n\n<p>where $t_{i,j}(p)$ is the trace of $Frob_p$ on $V_{i,j}$.</p>\n\n<p>Let me distinguish now between several different types of irreducible pieces:</p>\n\n<ul>\n<li>$V_{i, j}$ is a one-dimensional representation. Then $i$ must be even, and the trace of Frobenius on $V_{i, j}$ is $p^{i/2} \\chi(p)$ where $\\chi$ is a Dirichlet character. </li>\n<li>$V_{i, j}$ is two-dimensional and comes from a modular form. Then $t_{i,j}(p) = a_p(f)$, and $f$ must have weight $i+1$.</li>\n<li>$V_{i,j}$ is two-dimensional and doesn't come from a modular form. This can happen, but it's rare, and it's expected that all examples come from another kind of analytic object called a Maass wave form; in particular this forces $i$ to be even.</li>\n<li>$V_{i, j}$ has dimension $&gt; 2$. Then $V_{i, j}$ cannot be the Galois representation coming from a modular form, because these always have dimension 2.</li>\n</ul>\n\n<p>You seem to want your varieties to have $X(\\mathbf{F}_p)$ = (polynomial in $p$) + (coefficient of a modular form). From the above formulae, it's clear that this can only happen if all the $V_{i,j}$ have dimension 1 or 2; there is exactly one with dimension 2 and it comes from a modular form; and the one-dimensional pieces all come from the trivial Dirichlet character. This always happens for genus 1 curves, because the $H^0$ and $H^2$ are always 1-dimensional for a curve, and the genus condition forces the $H^1$ to be two-dimensional. </p>\n\n<p>However, once you step away from genus 1 curves, this is totally not the generic behaviour, and it will only occur for unusual and highly symmetric examples, such as the rigid Calabi-Yaus and extremal $K_3$ surfaces in the links you've posted. </p>\n" }, { "answer_id": 260185, "score": 2, "text": "<p>It seems that this holds as long as one counts with respect to the appropriate \"baseline\". In <a href=\"http://www.uni-math.gwdg.de/tschinkel/SS04/yui.pdf\" rel=\"nofollow noreferrer\">this paper</a> they give a class of examples. Let $X$ be a nice enough rigid Calabi-Yau threefold over $\\mathbb{Q}$, with Hodge numbers $h^{i,j}$. Then, if I understand them correctly (pg 7), they say that if we set</p>\n\n<p>$a_p := 1 + p^3 + (1+p)p h^{1,1} - \\#X(\\mathbb{F}_p)$</p>\n\n<p>then these are the Fourier coefficients (for almost all primes $p$) of a modular form of weight 4 and a certain level $N$. </p>\n\n<p>In general, it seems one needs to set the $a_p$ to be the trace of the Frobenius acting on etale cohomology in some degree $i$,</p>\n\n<p>$a_p = \\mbox{Tr}(\\mbox{Fr}_p : H^i(X) \\rightarrow H^i(X)$).</p>\n\n<p>In nice enough cases (such as the one above), one can interpret this trace concretely as a count of the points on $X$ mod p, with respect to an appropriate baseline.</p>\n" }, { "answer_id": 260232, "score": 22, "text": "<p>The correct setting for this construction turns out to be <em>projective</em> varieties, so let me suppose we have a smooth variety $X$ inside $\\mathbf{P}^N$, for some $N \\ge 1$, defined by the vanishing of some homogenous polynomials $F_1, \\dots, F_r$ in variables $x_0, \\dots, x_N$, with the $F_i$ having coefficients in $\\mathbf{Q}$. Actually, let me assume the $F_i$ have coefficients in $\\mathbf{Z}$, which is no loss since we can just multiply up. Then we can make sense of the reduction of $X$ modulo $p$; and we want to study the point counts $\\#X(\\mathbf{F}_p)$ as a function of $p$, possibly neglecting some finite set $\\Sigma$ containing all primes such that the reduction of $X$ mod $p$ is singular.</p>\n\n<p>Thanks to Grothendieck, Deligne, and others, we have a very powerful bunch of tools for analysing this problem. The setup is as follows. Choose your favourite prime $\\ell$. Then the theory of <strong>etale cohomology</strong> attaches to $X$ a bunch of finite-dimensional $\\mathbf{Q}_\\ell$-vector spaces $H^i_{\\mathrm{et}}(X_{\\overline{\\mathbf{Q}}}, \\mathbf{Q}_\\ell)$ (let me abbreviate this by $H^i_\\ell(X)$ to save typing). The dimension of $H^i_\\ell$ is the same as the $i$-th Betti number of the manifold $X(\\mathbb{C})$; but they encode much more data, because each $H^i_\\ell(X)$ is a representation of the Galois group $\\operatorname{Gal}(\\overline{\\mathbf{Q}} / \\mathbf{Q})$, unramified outside $\\Sigma \\cup \\{\\ell\\}$; so for every prime not in this set, and every $i$, we have a number $t_i(p)$, the trace of Frobenius at $p$ on $H^i_\\ell$, which turns out to be independent of $\\ell$.</p>\n\n<p><strong>Theorem</strong>: $\\#X(\\mathbf{F}_p)$ is the alternating sum $\\sum_{i=0}^{2 dim(X)} (-1)^i t_i(p)$.</p>\n\n<p>Now let's analyse $H^i_\\ell$ as a Galois representation. Representations of Galois groups needn't be direct sums of irreducibles, but we can replace $H^i_\\ell$ by its semisimplification, which does have this property and has the same trace as the original $H^i_\\ell$. This semisimplification will look like $V_{i, 1} + \\dots + V_{i, r_i}$ where $V_{i, j}$ are irreducible; and the $V_{i, j}$ all have motivic weight $i$, so the same $V$ can't appear for two different $i$'s. So we get a slightly finer decomposition</p>\n\n<p>$\\#X(\\mathbf{F}_p) = \\sum_{i=0}^{2 \\mathrm{dim} X} (-1)^i \\sum_{j=1}^{k_i} t_{i, j}(p)$</p>\n\n<p>where $t_{i,j}(p)$ is the trace of $Frob_p$ on $V_{i,j}$.</p>\n\n<p>Let me distinguish now between several different types of irreducible pieces:</p>\n\n<ul>\n<li>$V_{i, j}$ is a one-dimensional representation. Then $i$ must be even, and the trace of Frobenius on $V_{i, j}$ is $p^{i/2} \\chi(p)$ where $\\chi$ is a Dirichlet character. </li>\n<li>$V_{i, j}$ is two-dimensional and comes from a modular form. Then $t_{i,j}(p) = a_p(f)$, and $f$ must have weight $i+1$.</li>\n<li>$V_{i,j}$ is two-dimensional and doesn't come from a modular form. This can happen, but it's rare, and it's expected that all examples come from another kind of analytic object called a Maass wave form; in particular this forces $i$ to be even.</li>\n<li>$V_{i, j}$ has dimension $&gt; 2$. Then $V_{i, j}$ cannot be the Galois representation coming from a modular form, because these always have dimension 2.</li>\n</ul>\n\n<p>You seem to want your varieties to have $X(\\mathbf{F}_p)$ = (polynomial in $p$) + (coefficient of a modular form). From the above formulae, it's clear that this can only happen if all the $V_{i,j}$ have dimension 1 or 2; there is exactly one with dimension 2 and it comes from a modular form; and the one-dimensional pieces all come from the trivial Dirichlet character. This always happens for genus 1 curves, because the $H^0$ and $H^2$ are always 1-dimensional for a curve, and the genus condition forces the $H^1$ to be two-dimensional. </p>\n\n<p>However, once you step away from genus 1 curves, this is totally not the generic behaviour, and it will only occur for unusual and highly symmetric examples, such as the rigid Calabi-Yaus and extremal $K_3$ surfaces in the links you've posted. </p>\n" }, { "answer_id": 260185, "score": 2, "text": "<p>It seems that this holds as long as one counts with respect to the appropriate \"baseline\". In <a href=\"http://www.uni-math.gwdg.de/tschinkel/SS04/yui.pdf\" rel=\"nofollow noreferrer\">this paper</a> they give a class of examples. Let $X$ be a nice enough rigid Calabi-Yau threefold over $\\mathbb{Q}$, with Hodge numbers $h^{i,j}$. Then, if I understand them correctly (pg 7), they say that if we set</p>\n\n<p>$a_p := 1 + p^3 + (1+p)p h^{1,1} - \\#X(\\mathbb{F}_p)$</p>\n\n<p>then these are the Fourier coefficients (for almost all primes $p$) of a modular form of weight 4 and a certain level $N$. </p>\n\n<p>In general, it seems one needs to set the $a_p$ to be the trace of the Frobenius acting on etale cohomology in some degree $i$,</p>\n\n<p>$a_p = \\mbox{Tr}(\\mbox{Fr}_p : H^i(X) \\rightarrow H^i(X)$).</p>\n\n<p>In nice enough cases (such as the one above), one can interpret this trace concretely as a count of the points on $X$ mod p, with respect to an appropriate baseline.</p>\n" }, { "answer_id": 260232, "score": 22, "text": "<p>The correct setting for this construction turns out to be <em>projective</em> varieties, so let me suppose we have a smooth variety $X$ inside $\\mathbf{P}^N$, for some $N \\ge 1$, defined by the vanishing of some homogenous polynomials $F_1, \\dots, F_r$ in variables $x_0, \\dots, x_N$, with the $F_i$ having coefficients in $\\mathbf{Q}$. Actually, let me assume the $F_i$ have coefficients in $\\mathbf{Z}$, which is no loss since we can just multiply up. Then we can make sense of the reduction of $X$ modulo $p$; and we want to study the point counts $\\#X(\\mathbf{F}_p)$ as a function of $p$, possibly neglecting some finite set $\\Sigma$ containing all primes such that the reduction of $X$ mod $p$ is singular.</p>\n\n<p>Thanks to Grothendieck, Deligne, and others, we have a very powerful bunch of tools for analysing this problem. The setup is as follows. Choose your favourite prime $\\ell$. Then the theory of <strong>etale cohomology</strong> attaches to $X$ a bunch of finite-dimensional $\\mathbf{Q}_\\ell$-vector spaces $H^i_{\\mathrm{et}}(X_{\\overline{\\mathbf{Q}}}, \\mathbf{Q}_\\ell)$ (let me abbreviate this by $H^i_\\ell(X)$ to save typing). The dimension of $H^i_\\ell$ is the same as the $i$-th Betti number of the manifold $X(\\mathbb{C})$; but they encode much more data, because each $H^i_\\ell(X)$ is a representation of the Galois group $\\operatorname{Gal}(\\overline{\\mathbf{Q}} / \\mathbf{Q})$, unramified outside $\\Sigma \\cup \\{\\ell\\}$; so for every prime not in this set, and every $i$, we have a number $t_i(p)$, the trace of Frobenius at $p$ on $H^i_\\ell$, which turns out to be independent of $\\ell$.</p>\n\n<p><strong>Theorem</strong>: $\\#X(\\mathbf{F}_p)$ is the alternating sum $\\sum_{i=0}^{2 dim(X)} (-1)^i t_i(p)$.</p>\n\n<p>Now let's analyse $H^i_\\ell$ as a Galois representation. Representations of Galois groups needn't be direct sums of irreducibles, but we can replace $H^i_\\ell$ by its semisimplification, which does have this property and has the same trace as the original $H^i_\\ell$. This semisimplification will look like $V_{i, 1} + \\dots + V_{i, r_i}$ where $V_{i, j}$ are irreducible; and the $V_{i, j}$ all have motivic weight $i$, so the same $V$ can't appear for two different $i$'s. So we get a slightly finer decomposition</p>\n\n<p>$\\#X(\\mathbf{F}_p) = \\sum_{i=0}^{2 \\mathrm{dim} X} (-1)^i \\sum_{j=1}^{k_i} t_{i, j}(p)$</p>\n\n<p>where $t_{i,j}(p)$ is the trace of $Frob_p$ on $V_{i,j}$.</p>\n\n<p>Let me distinguish now between several different types of irreducible pieces:</p>\n\n<ul>\n<li>$V_{i, j}$ is a one-dimensional representation. Then $i$ must be even, and the trace of Frobenius on $V_{i, j}$ is $p^{i/2} \\chi(p)$ where $\\chi$ is a Dirichlet character. </li>\n<li>$V_{i, j}$ is two-dimensional and comes from a modular form. Then $t_{i,j}(p) = a_p(f)$, and $f$ must have weight $i+1$.</li>\n<li>$V_{i,j}$ is two-dimensional and doesn't come from a modular form. This can happen, but it's rare, and it's expected that all examples come from another kind of analytic object called a Maass wave form; in particular this forces $i$ to be even.</li>\n<li>$V_{i, j}$ has dimension $&gt; 2$. Then $V_{i, j}$ cannot be the Galois representation coming from a modular form, because these always have dimension 2.</li>\n</ul>\n\n<p>You seem to want your varieties to have $X(\\mathbf{F}_p)$ = (polynomial in $p$) + (coefficient of a modular form). From the above formulae, it's clear that this can only happen if all the $V_{i,j}$ have dimension 1 or 2; there is exactly one with dimension 2 and it comes from a modular form; and the one-dimensional pieces all come from the trivial Dirichlet character. This always happens for genus 1 curves, because the $H^0$ and $H^2$ are always 1-dimensional for a curve, and the genus condition forces the $H^1$ to be two-dimensional. </p>\n\n<p>However, once you step away from genus 1 curves, this is totally not the generic behaviour, and it will only occur for unusual and highly symmetric examples, such as the rigid Calabi-Yaus and extremal $K_3$ surfaces in the links you've posted. </p>\n" }, { "answer_id": 260185, "score": 2, "text": "<p>It seems that this holds as long as one counts with respect to the appropriate \"baseline\". In <a href=\"http://www.uni-math.gwdg.de/tschinkel/SS04/yui.pdf\" rel=\"nofollow noreferrer\">this paper</a> they give a class of examples. Let $X$ be a nice enough rigid Calabi-Yau threefold over $\\mathbb{Q}$, with Hodge numbers $h^{i,j}$. Then, if I understand them correctly (pg 7), they say that if we set</p>\n\n<p>$a_p := 1 + p^3 + (1+p)p h^{1,1} - \\#X(\\mathbb{F}_p)$</p>\n\n<p>then these are the Fourier coefficients (for almost all primes $p$) of a modular form of weight 4 and a certain level $N$. </p>\n\n<p>In general, it seems one needs to set the $a_p$ to be the trace of the Frobenius acting on etale cohomology in some degree $i$,</p>\n\n<p>$a_p = \\mbox{Tr}(\\mbox{Fr}_p : H^i(X) \\rightarrow H^i(X)$).</p>\n\n<p>In nice enough cases (such as the one above), one can interpret this trace concretely as a count of the points on $X$ mod p, with respect to an appropriate baseline.</p>\n" } ]
Question: Modular forms from counting points on algebraic varieties over a finite field Answer 1: <p>The correct setting for this construction turns out to be <em>projective</em> varieties, so let me suppose we have a smooth variety $X$ inside $\mathbf{P}^N$, for some $N \ge 1$, defined by the vanishing of some homogenous polynomials $F_1, \dots, F_r$ in variables $x_0, \dots, x_N$, with the $F_i$ having coefficients in $\mathbf{Q}$. Actually, let me assume the $F_i$ have coefficients in $\mathbf{Z}$, which is no loss since we can just multiply up. Then we can make sense of the reduction of $X$ modulo $p$; and we want to study the point counts $\#X(\mathbf{F}_p)$ as a function of $p$, possibly neglecting some finite set $\Sigma$ containing all primes such that the reduction of $X$ mod $p$ is singular.</p> <p>Thanks to Grothendieck, Deligne, and others, we have a very powerful bunch of tools for analysing this problem. The setup is as follows. Choose your favourite prime $\ell$. Then the theory of <strong>etale cohomology</strong> attaches to $X$ a bunch of finite-dimensional $\mathbf{Q}_\ell$-vector spaces $H^i_{\mathrm{et}}(X_{\overline{\mathbf{Q}}}, \mathbf{Q}_\ell)$ (let me abbreviate this by $H^i_\ell(X)$ to save typing). The dimension of $H^i_\ell$ is the same as the $i$-th Betti number of the manifold $X(\mathbb{C})$; but they encode much more data, because each $H^i_\ell(X)$ is a representation of the Galois group $\operatorname{Gal}(\overline{\mathbf{Q}} / \mathbf{Q})$, unramified outside $\Sigma \cup \{\ell\}$; so for every prime not in this set, and every $i$, we have a number $t_i(p)$, the trace of Frobenius at $p$ on $H^i_\ell$, which turns out to be independent of $\ell$.</p> <p><strong>Theorem</strong>: $\#X(\mathbf{F}_p)$ is the alternating sum $\sum_{i=0}^{2 dim(X)} (-1)^i t_i(p)$.</p> <p>Now let's analyse $H^i_\ell$ as a Galois representation. Representations of Galois groups needn't be direct sums of irreducibles, but we can replace $H^i_\ell$ by its semisimplification, which does have this property and has the same trace as the original $H^i_\ell$. This semisimplification will look like $V_{i, 1} + \dots + V_{i, r_i}$ where $V_{i, j}$ are irreducible; and the $V_{i, j}$ all have motivic weight $i$, so the same $V$ can't appear for two different $i$'s. So we get a slightly finer decomposition</p> <p>$\#X(\mathbf{F}_p) = \sum_{i=0}^{2 \mathrm{dim} X} (-1)^i \sum_{j=1}^{k_i} t_{i, j}(p)$</p> <p>where $t_{i,j}(p)$ is the trace of $Frob_p$ on $V_{i,j}$.</p> <p>Let me distinguish now between several different types of irreducible pieces:</p> <ul> <li>$V_{i, j}$ is a one-dimensional representation. Then $i$ must be even, and the trace of Frobenius on $V_{i, j}$ is $p^{i/2} \chi(p)$ where $\chi$ is a Dirichlet character. </li> <li>$V_{i, j}$ is two-dimensional and comes from a modular form. Then $t_{i,j}(p) = a_p(f)$, and $f$ must have weight $i+1$.</li> <li>$V_{i,j}$ is two-dimensional and doesn't come from a modular form. This can happen, but it's rare, and it's expected that all examples come from another kind of analytic object called a Maass wave form; in particular this forces $i$ to be even.</li> <li>$V_{i, j}$ has dimension $&gt; 2$. Then $V_{i, j}$ cannot be the Galois representation coming from a modular form, because these always have dimension 2.</li> </ul> <p>You seem to want your varieties to have $X(\mathbf{F}_p)$ = (polynomial in $p$) + (coefficient of a modular form). From the above formulae, it's clear that this can only happen if all the $V_{i,j}$ have dimension 1 or 2; there is exactly one with dimension 2 and it comes from a modular form; and the one-dimensional pieces all come from the trivial Dirichlet character. This always happens for genus 1 curves, because the $H^0$ and $H^2$ are always 1-dimensional for a curve, and the genus condition forces the $H^1$ to be two-dimensional. </p> <p>However, once you step away from genus 1 curves, this is totally not the generic behaviour, and it will only occur for unusual and highly symmetric examples, such as the rigid Calabi-Yaus and extremal $K_3$ surfaces in the links you've posted. </p> Answer 2: <p>It seems that this holds as long as one counts with respect to the appropriate "baseline". In <a href="http://www.uni-math.gwdg.de/tschinkel/SS04/yui.pdf" rel="nofollow noreferrer">this paper</a> they give a class of examples. Let $X$ be a nice enough rigid Calabi-Yau threefold over $\mathbb{Q}$, with Hodge numbers $h^{i,j}$. Then, if I understand them correctly (pg 7), they say that if we set</p> <p>$a_p := 1 + p^3 + (1+p)p h^{1,1} - \#X(\mathbb{F}_p)$</p> <p>then these are the Fourier coefficients (for almost all primes $p$) of a modular form of weight 4 and a certain level $N$. </p> <p>In general, it seems one needs to set the $a_p$ to be the trace of the Frobenius acting on etale cohomology in some degree $i$,</p> <p>$a_p = \mbox{Tr}(\mbox{Fr}_p : H^i(X) \rightarrow H^i(X)$).</p> <p>In nice enough cases (such as the one above), one can interpret this trace concretely as a count of the points on $X$ mod p, with respect to an appropriate baseline.</p> Answer 3: <p>The correct setting for this construction turns out to be <em>projective</em> varieties, so let me suppose we have a smooth variety $X$ inside $\mathbf{P}^N$, for some $N \ge 1$, defined by the vanishing of some homogenous polynomials $F_1, \dots, F_r$ in variables $x_0, \dots, x_N$, with the $F_i$ having coefficients in $\mathbf{Q}$. Actually, let me assume the $F_i$ have coefficients in $\mathbf{Z}$, which is no loss since we can just multiply up. Then we can make sense of the reduction of $X$ modulo $p$; and we want to study the point counts $\#X(\mathbf{F}_p)$ as a function of $p$, possibly neglecting some finite set $\Sigma$ containing all primes such that the reduction of $X$ mod $p$ is singular.</p> <p>Thanks to Grothendieck, Deligne, and others, we have a very powerful bunch of tools for analysing this problem. The setup is as follows. Choose your favourite prime $\ell$. Then the theory of <strong>etale cohomology</strong> attaches to $X$ a bunch of finite-dimensional $\mathbf{Q}_\ell$-vector spaces $H^i_{\mathrm{et}}(X_{\overline{\mathbf{Q}}}, \mathbf{Q}_\ell)$ (let me abbreviate this by $H^i_\ell(X)$ to save typing). The dimension of $H^i_\ell$ is the same as the $i$-th Betti number of the manifold $X(\mathbb{C})$; but they encode much more data, because each $H^i_\ell(X)$ is a representation of the Galois group $\operatorname{Gal}(\overline{\mathbf{Q}} / \mathbf{Q})$, unramified outside $\Sigma \cup \{\ell\}$; so for every prime not in this set, and every $i$, we have a number $t_i(p)$, the trace of Frobenius at $p$ on $H^i_\ell$, which turns out to be independent of $\ell$.</p> <p><strong>Theorem</strong>: $\#X(\mathbf{F}_p)$ is the alternating sum $\sum_{i=0}^{2 dim(X)} (-1)^i t_i(p)$.</p> <p>Now let's analyse $H^i_\ell$ as a Galois representation. Representations of Galois groups needn't be direct sums of irreducibles, but we can replace $H^i_\ell$ by its semisimplification, which does have this property and has the same trace as the original $H^i_\ell$. This semisimplification will look like $V_{i, 1} + \dots + V_{i, r_i}$ where $V_{i, j}$ are irreducible; and the $V_{i, j}$ all have motivic weight $i$, so the same $V$ can't appear for two different $i$'s. So we get a slightly finer decomposition</p> <p>$\#X(\mathbf{F}_p) = \sum_{i=0}^{2 \mathrm{dim} X} (-1)^i \sum_{j=1}^{k_i} t_{i, j}(p)$</p> <p>where $t_{i,j}(p)$ is the trace of $Frob_p$ on $V_{i,j}$.</p> <p>Let me distinguish now between several different types of irreducible pieces:</p> <ul> <li>$V_{i, j}$ is a one-dimensional representation. Then $i$ must be even, and the trace of Frobenius on $V_{i, j}$ is $p^{i/2} \chi(p)$ where $\chi$ is a Dirichlet character. </li> <li>$V_{i, j}$ is two-dimensional and comes from a modular form. Then $t_{i,j}(p) = a_p(f)$, and $f$ must have weight $i+1$.</li> <li>$V_{i,j}$ is two-dimensional and doesn't come from a modular form. This can happen, but it's rare, and it's expected that all examples come from another kind of analytic object called a Maass wave form; in particular this forces $i$ to be even.</li> <li>$V_{i, j}$ has dimension $&gt; 2$. Then $V_{i, j}$ cannot be the Galois representation coming from a modular form, because these always have dimension 2.</li> </ul> <p>You seem to want your varieties to have $X(\mathbf{F}_p)$ = (polynomial in $p$) + (coefficient of a modular form). From the above formulae, it's clear that this can only happen if all the $V_{i,j}$ have dimension 1 or 2; there is exactly one with dimension 2 and it comes from a modular form; and the one-dimensional pieces all come from the trivial Dirichlet character. This always happens for genus 1 curves, because the $H^0$ and $H^2$ are always 1-dimensional for a curve, and the genus condition forces the $H^1$ to be two-dimensional. </p> <p>However, once you step away from genus 1 curves, this is totally not the generic behaviour, and it will only occur for unusual and highly symmetric examples, such as the rigid Calabi-Yaus and extremal $K_3$ surfaces in the links you've posted. </p> Answer 4: <p>It seems that this holds as long as one counts with respect to the appropriate "baseline". In <a href="http://www.uni-math.gwdg.de/tschinkel/SS04/yui.pdf" rel="nofollow noreferrer">this paper</a> they give a class of examples. Let $X$ be a nice enough rigid Calabi-Yau threefold over $\mathbb{Q}$, with Hodge numbers $h^{i,j}$. Then, if I understand them correctly (pg 7), they say that if we set</p> <p>$a_p := 1 + p^3 + (1+p)p h^{1,1} - \#X(\mathbb{F}_p)$</p> <p>then these are the Fourier coefficients (for almost all primes $p$) of a modular form of weight 4 and a certain level $N$. </p> <p>In general, it seems one needs to set the $a_p$ to be the trace of the Frobenius acting on etale cohomology in some degree $i$,</p> <p>$a_p = \mbox{Tr}(\mbox{Fr}_p : H^i(X) \rightarrow H^i(X)$).</p> <p>In nice enough cases (such as the one above), one can interpret this trace concretely as a count of the points on $X$ mod p, with respect to an appropriate baseline.</p> Answer 5: <p>The correct setting for this construction turns out to be <em>projective</em> varieties, so let me suppose we have a smooth variety $X$ inside $\mathbf{P}^N$, for some $N \ge 1$, defined by the vanishing of some homogenous polynomials $F_1, \dots, F_r$ in variables $x_0, \dots, x_N$, with the $F_i$ having coefficients in $\mathbf{Q}$. Actually, let me assume the $F_i$ have coefficients in $\mathbf{Z}$, which is no loss since we can just multiply up. Then we can make sense of the reduction of $X$ modulo $p$; and we want to study the point counts $\#X(\mathbf{F}_p)$ as a function of $p$, possibly neglecting some finite set $\Sigma$ containing all primes such that the reduction of $X$ mod $p$ is singular.</p> <p>Thanks to Grothendieck, Deligne, and others, we have a very powerful bunch of tools for analysing this problem. The setup is as follows. Choose your favourite prime $\ell$. Then the theory of <strong>etale cohomology</strong> attaches to $X$ a bunch of finite-dimensional $\mathbf{Q}_\ell$-vector spaces $H^i_{\mathrm{et}}(X_{\overline{\mathbf{Q}}}, \mathbf{Q}_\ell)$ (let me abbreviate this by $H^i_\ell(X)$ to save typing). The dimension of $H^i_\ell$ is the same as the $i$-th Betti number of the manifold $X(\mathbb{C})$; but they encode much more data, because each $H^i_\ell(X)$ is a representation of the Galois group $\operatorname{Gal}(\overline{\mathbf{Q}} / \mathbf{Q})$, unramified outside $\Sigma \cup \{\ell\}$; so for every prime not in this set, and every $i$, we have a number $t_i(p)$, the trace of Frobenius at $p$ on $H^i_\ell$, which turns out to be independent of $\ell$.</p> <p><strong>Theorem</strong>: $\#X(\mathbf{F}_p)$ is the alternating sum $\sum_{i=0}^{2 dim(X)} (-1)^i t_i(p)$.</p> <p>Now let's analyse $H^i_\ell$ as a Galois representation. Representations of Galois groups needn't be direct sums of irreducibles, but we can replace $H^i_\ell$ by its semisimplification, which does have this property and has the same trace as the original $H^i_\ell$. This semisimplification will look like $V_{i, 1} + \dots + V_{i, r_i}$ where $V_{i, j}$ are irreducible; and the $V_{i, j}$ all have motivic weight $i$, so the same $V$ can't appear for two different $i$'s. So we get a slightly finer decomposition</p> <p>$\#X(\mathbf{F}_p) = \sum_{i=0}^{2 \mathrm{dim} X} (-1)^i \sum_{j=1}^{k_i} t_{i, j}(p)$</p> <p>where $t_{i,j}(p)$ is the trace of $Frob_p$ on $V_{i,j}$.</p> <p>Let me distinguish now between several different types of irreducible pieces:</p> <ul> <li>$V_{i, j}$ is a one-dimensional representation. Then $i$ must be even, and the trace of Frobenius on $V_{i, j}$ is $p^{i/2} \chi(p)$ where $\chi$ is a Dirichlet character. </li> <li>$V_{i, j}$ is two-dimensional and comes from a modular form. Then $t_{i,j}(p) = a_p(f)$, and $f$ must have weight $i+1$.</li> <li>$V_{i,j}$ is two-dimensional and doesn't come from a modular form. This can happen, but it's rare, and it's expected that all examples come from another kind of analytic object called a Maass wave form; in particular this forces $i$ to be even.</li> <li>$V_{i, j}$ has dimension $&gt; 2$. Then $V_{i, j}$ cannot be the Galois representation coming from a modular form, because these always have dimension 2.</li> </ul> <p>You seem to want your varieties to have $X(\mathbf{F}_p)$ = (polynomial in $p$) + (coefficient of a modular form). From the above formulae, it's clear that this can only happen if all the $V_{i,j}$ have dimension 1 or 2; there is exactly one with dimension 2 and it comes from a modular form; and the one-dimensional pieces all come from the trivial Dirichlet character. This always happens for genus 1 curves, because the $H^0$ and $H^2$ are always 1-dimensional for a curve, and the genus condition forces the $H^1$ to be two-dimensional. </p> <p>However, once you step away from genus 1 curves, this is totally not the generic behaviour, and it will only occur for unusual and highly symmetric examples, such as the rigid Calabi-Yaus and extremal $K_3$ surfaces in the links you've posted. </p> Answer 6: <p>It seems that this holds as long as one counts with respect to the appropriate "baseline". In <a href="http://www.uni-math.gwdg.de/tschinkel/SS04/yui.pdf" rel="nofollow noreferrer">this paper</a> they give a class of examples. Let $X$ be a nice enough rigid Calabi-Yau threefold over $\mathbb{Q}$, with Hodge numbers $h^{i,j}$. Then, if I understand them correctly (pg 7), they say that if we set</p> <p>$a_p := 1 + p^3 + (1+p)p h^{1,1} - \#X(\mathbb{F}_p)$</p> <p>then these are the Fourier coefficients (for almost all primes $p$) of a modular form of weight 4 and a certain level $N$. </p> <p>In general, it seems one needs to set the $a_p$ to be the trace of the Frobenius acting on etale cohomology in some degree $i$,</p> <p>$a_p = \mbox{Tr}(\mbox{Fr}_p : H^i(X) \rightarrow H^i(X)$).</p> <p>In nice enough cases (such as the one above), one can interpret this trace concretely as a count of the points on $X$ mod p, with respect to an appropriate baseline.</p>
Does the class of Hausdorff spaces have a shared &quot;Coordinate space&quot;?
260,329
15
https://mathoverflow.net/questions/260329/does-the-class-of-hausdorff-spaces-have-a-shared-coordinate-space
<p>Given topological spaces $X$ and $C$ we call $C$ a <strong>coordinate space</strong> for $X$ to mean that every open set $U \subset X$ is of the form $f^{-1}(V)$ for some open $V \subset C$ and continuous $f \colon X \to C$. </p> <p>More generally if $\mathscr X$ is a class of spaces we call $C$ a <strong>coordinate space</strong> for $\mathscr X$ to mean it is a coordinate space for every member of $\mathscr X$.</p> <p>Two familiar examples of coordinate spaces:</p> <ol> <li><p>The Seirpinski space $\{0,1\}$ with exactly one closed point and one open point is a coordinate space for the class of all topological spaces.</p></li> <li><p>The closed unit interval is a coordinate space for the class of all completely-regular spaces. Moreover the closed unit interval is itself completely-regular.</p></li> </ol> <p>But what about the class of all Hausdorff spaces? Obviously the Seirpinski space is still a coordinate space here. But what if we demand the coordinate space itself be Hausdorff -- as with example 2? </p> <p>I would reckon such a space does not exist but have no idea how to prove that?</p> <p>Does this notion already have a name and has the question been addressed before?</p>
[ { "answer_id": 260349, "score": 15, "text": "<p>First of all, let me point out that $[0,1]$ is not a coordinate space for the class of completely regular spaces.</p>\n\n<p>The definition of completely regular spaces says that for any closed $C \\subset X$ and any point $p \\notin C$, there is a continuous $[0,1]$-valued function (say $f$) mapping $p$ to $1$ and mapping every point in $C$ to $0$. It does <em>not</em> say that $f^{-1}(0,1] = X-C$. In other words, $f$ does map $C$ to $0$, but it may also map some other points to $0$ as well.</p>\n\n<p>For example, it can be shown that any continuous function $f: \\omega_1 \\rightarrow [0,1]$ is eventually constant. It follows that $[0,1]$ is not a coordinate space for $\\omega_1+1$: the open set $\\omega_1$ is never equal to $f^{-1}(U)$ for an open $U \\subseteq [0,1]$ and continuous $f: \\omega_1+1 \\rightarrow [0,1]$, because $f$ maps a tail of $\\omega_1$ to a single point $c$, and must (by continuity) map a tail of $\\omega_1+1$ to $c$. So if $f^{-1}(U)$ contains $\\omega_1$, it contains $\\omega_1+1$.</p>\n\n<p>A similar argument shows</p>\n\n<blockquote>\n <p>There is no Hausdorff coordinate space for the class of Hausdorff spaces.</p>\n</blockquote>\n\n<p>To prove this, suppose $C$ is such a space. Let $\\kappa$ be a regular cardinal bigger than $|C|$. Any function $\\kappa \\rightarrow C$ must map an unbounded subset of $\\kappa$ to a single point (there's no topology involved in this statement -- it's just because $\\kappa$ is regular and bigger than $|C|$). Now consider the ordinal space $\\kappa+1$. As above, the open set $\\kappa$ is never equal to $f^{-1}(U)$ for an open $U \\subseteq C$ and continuous $f: \\kappa+1 \\rightarrow C$.</p>\n" }, { "answer_id": 260367, "score": 3, "text": "<p>A classical notion was considered and solved in an elegant paper by Stanisław Mrówka: (1) instead of arbitrary open subspaces of $\\ X\\ $ one simply considers just X itself; and (2) instead of a single space $\\ C,\\ $ one considered the class of all powers $\\ C^S.\\ $ Then Stanisław Mrówka answered that this kind of embeddability problem is characterized by the property of separating points from closed subsets of $\\ X\\ $ by functions into $\\ C^F,\\ $ where this time the exponents $\\ f\\ $ are just finite sets--a very elegant conceptual theorem.</p>\n\n<p><strong>Acknowledgment</strong>--thank you, Will Brian, for pointing to my earlier error.</p>\n" }, { "answer_id": 260349, "score": 15, "text": "<p>First of all, let me point out that $[0,1]$ is not a coordinate space for the class of completely regular spaces.</p>\n\n<p>The definition of completely regular spaces says that for any closed $C \\subset X$ and any point $p \\notin C$, there is a continuous $[0,1]$-valued function (say $f$) mapping $p$ to $1$ and mapping every point in $C$ to $0$. It does <em>not</em> say that $f^{-1}(0,1] = X-C$. In other words, $f$ does map $C$ to $0$, but it may also map some other points to $0$ as well.</p>\n\n<p>For example, it can be shown that any continuous function $f: \\omega_1 \\rightarrow [0,1]$ is eventually constant. It follows that $[0,1]$ is not a coordinate space for $\\omega_1+1$: the open set $\\omega_1$ is never equal to $f^{-1}(U)$ for an open $U \\subseteq [0,1]$ and continuous $f: \\omega_1+1 \\rightarrow [0,1]$, because $f$ maps a tail of $\\omega_1$ to a single point $c$, and must (by continuity) map a tail of $\\omega_1+1$ to $c$. So if $f^{-1}(U)$ contains $\\omega_1$, it contains $\\omega_1+1$.</p>\n\n<p>A similar argument shows</p>\n\n<blockquote>\n <p>There is no Hausdorff coordinate space for the class of Hausdorff spaces.</p>\n</blockquote>\n\n<p>To prove this, suppose $C$ is such a space. Let $\\kappa$ be a regular cardinal bigger than $|C|$. Any function $\\kappa \\rightarrow C$ must map an unbounded subset of $\\kappa$ to a single point (there's no topology involved in this statement -- it's just because $\\kappa$ is regular and bigger than $|C|$). Now consider the ordinal space $\\kappa+1$. As above, the open set $\\kappa$ is never equal to $f^{-1}(U)$ for an open $U \\subseteq C$ and continuous $f: \\kappa+1 \\rightarrow C$.</p>\n" }, { "answer_id": 260367, "score": 3, "text": "<p>A classical notion was considered and solved in an elegant paper by Stanisław Mrówka: (1) instead of arbitrary open subspaces of $\\ X\\ $ one simply considers just X itself; and (2) instead of a single space $\\ C,\\ $ one considered the class of all powers $\\ C^S.\\ $ Then Stanisław Mrówka answered that this kind of embeddability problem is characterized by the property of separating points from closed subsets of $\\ X\\ $ by functions into $\\ C^F,\\ $ where this time the exponents $\\ f\\ $ are just finite sets--a very elegant conceptual theorem.</p>\n\n<p><strong>Acknowledgment</strong>--thank you, Will Brian, for pointing to my earlier error.</p>\n" }, { "answer_id": 260349, "score": 15, "text": "<p>First of all, let me point out that $[0,1]$ is not a coordinate space for the class of completely regular spaces.</p>\n\n<p>The definition of completely regular spaces says that for any closed $C \\subset X$ and any point $p \\notin C$, there is a continuous $[0,1]$-valued function (say $f$) mapping $p$ to $1$ and mapping every point in $C$ to $0$. It does <em>not</em> say that $f^{-1}(0,1] = X-C$. In other words, $f$ does map $C$ to $0$, but it may also map some other points to $0$ as well.</p>\n\n<p>For example, it can be shown that any continuous function $f: \\omega_1 \\rightarrow [0,1]$ is eventually constant. It follows that $[0,1]$ is not a coordinate space for $\\omega_1+1$: the open set $\\omega_1$ is never equal to $f^{-1}(U)$ for an open $U \\subseteq [0,1]$ and continuous $f: \\omega_1+1 \\rightarrow [0,1]$, because $f$ maps a tail of $\\omega_1$ to a single point $c$, and must (by continuity) map a tail of $\\omega_1+1$ to $c$. So if $f^{-1}(U)$ contains $\\omega_1$, it contains $\\omega_1+1$.</p>\n\n<p>A similar argument shows</p>\n\n<blockquote>\n <p>There is no Hausdorff coordinate space for the class of Hausdorff spaces.</p>\n</blockquote>\n\n<p>To prove this, suppose $C$ is such a space. Let $\\kappa$ be a regular cardinal bigger than $|C|$. Any function $\\kappa \\rightarrow C$ must map an unbounded subset of $\\kappa$ to a single point (there's no topology involved in this statement -- it's just because $\\kappa$ is regular and bigger than $|C|$). Now consider the ordinal space $\\kappa+1$. As above, the open set $\\kappa$ is never equal to $f^{-1}(U)$ for an open $U \\subseteq C$ and continuous $f: \\kappa+1 \\rightarrow C$.</p>\n" }, { "answer_id": 260367, "score": 3, "text": "<p>A classical notion was considered and solved in an elegant paper by Stanisław Mrówka: (1) instead of arbitrary open subspaces of $\\ X\\ $ one simply considers just X itself; and (2) instead of a single space $\\ C,\\ $ one considered the class of all powers $\\ C^S.\\ $ Then Stanisław Mrówka answered that this kind of embeddability problem is characterized by the property of separating points from closed subsets of $\\ X\\ $ by functions into $\\ C^F,\\ $ where this time the exponents $\\ f\\ $ are just finite sets--a very elegant conceptual theorem.</p>\n\n<p><strong>Acknowledgment</strong>--thank you, Will Brian, for pointing to my earlier error.</p>\n" } ]
Question: Does the class of Hausdorff spaces have a shared &quot;Coordinate space&quot;? Answer 1: <p>First of all, let me point out that $[0,1]$ is not a coordinate space for the class of completely regular spaces.</p> <p>The definition of completely regular spaces says that for any closed $C \subset X$ and any point $p \notin C$, there is a continuous $[0,1]$-valued function (say $f$) mapping $p$ to $1$ and mapping every point in $C$ to $0$. It does <em>not</em> say that $f^{-1}(0,1] = X-C$. In other words, $f$ does map $C$ to $0$, but it may also map some other points to $0$ as well.</p> <p>For example, it can be shown that any continuous function $f: \omega_1 \rightarrow [0,1]$ is eventually constant. It follows that $[0,1]$ is not a coordinate space for $\omega_1+1$: the open set $\omega_1$ is never equal to $f^{-1}(U)$ for an open $U \subseteq [0,1]$ and continuous $f: \omega_1+1 \rightarrow [0,1]$, because $f$ maps a tail of $\omega_1$ to a single point $c$, and must (by continuity) map a tail of $\omega_1+1$ to $c$. So if $f^{-1}(U)$ contains $\omega_1$, it contains $\omega_1+1$.</p> <p>A similar argument shows</p> <blockquote> <p>There is no Hausdorff coordinate space for the class of Hausdorff spaces.</p> </blockquote> <p>To prove this, suppose $C$ is such a space. Let $\kappa$ be a regular cardinal bigger than $|C|$. Any function $\kappa \rightarrow C$ must map an unbounded subset of $\kappa$ to a single point (there's no topology involved in this statement -- it's just because $\kappa$ is regular and bigger than $|C|$). Now consider the ordinal space $\kappa+1$. As above, the open set $\kappa$ is never equal to $f^{-1}(U)$ for an open $U \subseteq C$ and continuous $f: \kappa+1 \rightarrow C$.</p> Answer 2: <p>A classical notion was considered and solved in an elegant paper by Stanisław Mrówka: (1) instead of arbitrary open subspaces of $\ X\ $ one simply considers just X itself; and (2) instead of a single space $\ C,\ $ one considered the class of all powers $\ C^S.\ $ Then Stanisław Mrówka answered that this kind of embeddability problem is characterized by the property of separating points from closed subsets of $\ X\ $ by functions into $\ C^F,\ $ where this time the exponents $\ f\ $ are just finite sets--a very elegant conceptual theorem.</p> <p><strong>Acknowledgment</strong>--thank you, Will Brian, for pointing to my earlier error.</p> Answer 3: <p>First of all, let me point out that $[0,1]$ is not a coordinate space for the class of completely regular spaces.</p> <p>The definition of completely regular spaces says that for any closed $C \subset X$ and any point $p \notin C$, there is a continuous $[0,1]$-valued function (say $f$) mapping $p$ to $1$ and mapping every point in $C$ to $0$. It does <em>not</em> say that $f^{-1}(0,1] = X-C$. In other words, $f$ does map $C$ to $0$, but it may also map some other points to $0$ as well.</p> <p>For example, it can be shown that any continuous function $f: \omega_1 \rightarrow [0,1]$ is eventually constant. It follows that $[0,1]$ is not a coordinate space for $\omega_1+1$: the open set $\omega_1$ is never equal to $f^{-1}(U)$ for an open $U \subseteq [0,1]$ and continuous $f: \omega_1+1 \rightarrow [0,1]$, because $f$ maps a tail of $\omega_1$ to a single point $c$, and must (by continuity) map a tail of $\omega_1+1$ to $c$. So if $f^{-1}(U)$ contains $\omega_1$, it contains $\omega_1+1$.</p> <p>A similar argument shows</p> <blockquote> <p>There is no Hausdorff coordinate space for the class of Hausdorff spaces.</p> </blockquote> <p>To prove this, suppose $C$ is such a space. Let $\kappa$ be a regular cardinal bigger than $|C|$. Any function $\kappa \rightarrow C$ must map an unbounded subset of $\kappa$ to a single point (there's no topology involved in this statement -- it's just because $\kappa$ is regular and bigger than $|C|$). Now consider the ordinal space $\kappa+1$. As above, the open set $\kappa$ is never equal to $f^{-1}(U)$ for an open $U \subseteq C$ and continuous $f: \kappa+1 \rightarrow C$.</p> Answer 4: <p>A classical notion was considered and solved in an elegant paper by Stanisław Mrówka: (1) instead of arbitrary open subspaces of $\ X\ $ one simply considers just X itself; and (2) instead of a single space $\ C,\ $ one considered the class of all powers $\ C^S.\ $ Then Stanisław Mrówka answered that this kind of embeddability problem is characterized by the property of separating points from closed subsets of $\ X\ $ by functions into $\ C^F,\ $ where this time the exponents $\ f\ $ are just finite sets--a very elegant conceptual theorem.</p> <p><strong>Acknowledgment</strong>--thank you, Will Brian, for pointing to my earlier error.</p> Answer 5: <p>First of all, let me point out that $[0,1]$ is not a coordinate space for the class of completely regular spaces.</p> <p>The definition of completely regular spaces says that for any closed $C \subset X$ and any point $p \notin C$, there is a continuous $[0,1]$-valued function (say $f$) mapping $p$ to $1$ and mapping every point in $C$ to $0$. It does <em>not</em> say that $f^{-1}(0,1] = X-C$. In other words, $f$ does map $C$ to $0$, but it may also map some other points to $0$ as well.</p> <p>For example, it can be shown that any continuous function $f: \omega_1 \rightarrow [0,1]$ is eventually constant. It follows that $[0,1]$ is not a coordinate space for $\omega_1+1$: the open set $\omega_1$ is never equal to $f^{-1}(U)$ for an open $U \subseteq [0,1]$ and continuous $f: \omega_1+1 \rightarrow [0,1]$, because $f$ maps a tail of $\omega_1$ to a single point $c$, and must (by continuity) map a tail of $\omega_1+1$ to $c$. So if $f^{-1}(U)$ contains $\omega_1$, it contains $\omega_1+1$.</p> <p>A similar argument shows</p> <blockquote> <p>There is no Hausdorff coordinate space for the class of Hausdorff spaces.</p> </blockquote> <p>To prove this, suppose $C$ is such a space. Let $\kappa$ be a regular cardinal bigger than $|C|$. Any function $\kappa \rightarrow C$ must map an unbounded subset of $\kappa$ to a single point (there's no topology involved in this statement -- it's just because $\kappa$ is regular and bigger than $|C|$). Now consider the ordinal space $\kappa+1$. As above, the open set $\kappa$ is never equal to $f^{-1}(U)$ for an open $U \subseteq C$ and continuous $f: \kappa+1 \rightarrow C$.</p> Answer 6: <p>A classical notion was considered and solved in an elegant paper by Stanisław Mrówka: (1) instead of arbitrary open subspaces of $\ X\ $ one simply considers just X itself; and (2) instead of a single space $\ C,\ $ one considered the class of all powers $\ C^S.\ $ Then Stanisław Mrówka answered that this kind of embeddability problem is characterized by the property of separating points from closed subsets of $\ X\ $ by functions into $\ C^F,\ $ where this time the exponents $\ f\ $ are just finite sets--a very elegant conceptual theorem.</p> <p><strong>Acknowledgment</strong>--thank you, Will Brian, for pointing to my earlier error.</p>
&quot;Correct&quot; definition of stratified spaces and reference for constructible sheaves?
258,562
14
https://mathoverflow.net/questions/258562/correct-definition-of-stratified-spaces-and-reference-for-constructible-sheave
<p>It seems that the theory of constructible sheaves (in particular anything that goes into proving that they form an abelian category) requires some technical statements about existence of certain stratifications and yet at the same time the notion of a &quot;stratified space&quot; seems to have several inequivalent definitions in the literature.</p> <blockquote> <ol> <li><p>Is there (as of today) an agreement among specialists in the field for <strong>what is the &quot;correct&quot; definition of a &quot;stratified space&quot;?</strong></p> </li> <li><p><strong>What would be a good reference to read about stratified spaces, constructible sheaves and their six functors?</strong> (in particular a complete construction of the abelian and derived categories of constructible sheaves). (Perhaps with a little bit of stratified Morse theory too).</p> </li> </ol> </blockquote> <p>My end goal is to have a concrete understanding of perverse sheaves on algebraic varieties and their stratifications. I don't want to get bogged down in the technicalities of stratified spaces but I would like to know how stratifications look and how to construct them in this context. However if there's a &quot;natural theory&quot; (and in particular a natural category) of stratified spaces I'd love to read about it.</p>
[ { "answer_id": 258565, "score": 8, "text": "<p>Consult the book <em>\"Sheaves on manifolds\"</em> by Kashiwara and Schapira. It's a hard nut to crack, but it is the most efficient presentation I've seen. </p>\n\n<p>In Chapter 8, they work with stratifications whose strata satisfy the <em>$\\mu$-regularity condition</em>. This is equivalent to the Verdier regularity condition which in turn is a bit stronger than the Whitney condition b. </p>\n\n<p>One can show that any subanalytic set admits stratifications satisfying the $\\mu$-regularity conditions. In particular, this condition implies that such stratified spaces admit (many) stratified Morse functions.</p>\n" }, { "answer_id": 259215, "score": 6, "text": "<p>A bit more reader friendly than Kashiwara-Schapira is Borel's Intersection Cohomology. It doesn't treat perverse sheaves, but it has a good overview of stratified spaces in the first few chapter and develops constructible sheaves and the six functors in Chapter V. You might also try Dimca's Sheaves in Topology, which does talk about perverse sheaves. Neither of these present the most general possible development, but they're pretty accessible.</p>\n\n<p>I'll also plug the discussion of stratified spaces in my book-in-progress, though I take a more topological approach that's not particularly oriented toward sheaves. Here's a link: <a href=\"http://faculty.tcu.edu/gfriedman/IHbook.pdf\" rel=\"nofollow noreferrer\">http://faculty.tcu.edu/gfriedman/IHbook.pdf</a></p>\n" }, { "answer_id": 258565, "score": 8, "text": "<p>Consult the book <em>\"Sheaves on manifolds\"</em> by Kashiwara and Schapira. It's a hard nut to crack, but it is the most efficient presentation I've seen. </p>\n\n<p>In Chapter 8, they work with stratifications whose strata satisfy the <em>$\\mu$-regularity condition</em>. This is equivalent to the Verdier regularity condition which in turn is a bit stronger than the Whitney condition b. </p>\n\n<p>One can show that any subanalytic set admits stratifications satisfying the $\\mu$-regularity conditions. In particular, this condition implies that such stratified spaces admit (many) stratified Morse functions.</p>\n" }, { "answer_id": 259215, "score": 6, "text": "<p>A bit more reader friendly than Kashiwara-Schapira is Borel's Intersection Cohomology. It doesn't treat perverse sheaves, but it has a good overview of stratified spaces in the first few chapter and develops constructible sheaves and the six functors in Chapter V. You might also try Dimca's Sheaves in Topology, which does talk about perverse sheaves. Neither of these present the most general possible development, but they're pretty accessible.</p>\n\n<p>I'll also plug the discussion of stratified spaces in my book-in-progress, though I take a more topological approach that's not particularly oriented toward sheaves. Here's a link: <a href=\"http://faculty.tcu.edu/gfriedman/IHbook.pdf\" rel=\"nofollow noreferrer\">http://faculty.tcu.edu/gfriedman/IHbook.pdf</a></p>\n" }, { "answer_id": 258565, "score": 8, "text": "<p>Consult the book <em>\"Sheaves on manifolds\"</em> by Kashiwara and Schapira. It's a hard nut to crack, but it is the most efficient presentation I've seen. </p>\n\n<p>In Chapter 8, they work with stratifications whose strata satisfy the <em>$\\mu$-regularity condition</em>. This is equivalent to the Verdier regularity condition which in turn is a bit stronger than the Whitney condition b. </p>\n\n<p>One can show that any subanalytic set admits stratifications satisfying the $\\mu$-regularity conditions. In particular, this condition implies that such stratified spaces admit (many) stratified Morse functions.</p>\n" }, { "answer_id": 259215, "score": 6, "text": "<p>A bit more reader friendly than Kashiwara-Schapira is Borel's Intersection Cohomology. It doesn't treat perverse sheaves, but it has a good overview of stratified spaces in the first few chapter and develops constructible sheaves and the six functors in Chapter V. You might also try Dimca's Sheaves in Topology, which does talk about perverse sheaves. Neither of these present the most general possible development, but they're pretty accessible.</p>\n\n<p>I'll also plug the discussion of stratified spaces in my book-in-progress, though I take a more topological approach that's not particularly oriented toward sheaves. Here's a link: <a href=\"http://faculty.tcu.edu/gfriedman/IHbook.pdf\" rel=\"nofollow noreferrer\">http://faculty.tcu.edu/gfriedman/IHbook.pdf</a></p>\n" } ]
Question: &quot;Correct&quot; definition of stratified spaces and reference for constructible sheaves? Answer 1: <p>Consult the book <em>"Sheaves on manifolds"</em> by Kashiwara and Schapira. It's a hard nut to crack, but it is the most efficient presentation I've seen. </p> <p>In Chapter 8, they work with stratifications whose strata satisfy the <em>$\mu$-regularity condition</em>. This is equivalent to the Verdier regularity condition which in turn is a bit stronger than the Whitney condition b. </p> <p>One can show that any subanalytic set admits stratifications satisfying the $\mu$-regularity conditions. In particular, this condition implies that such stratified spaces admit (many) stratified Morse functions.</p> Answer 2: <p>A bit more reader friendly than Kashiwara-Schapira is Borel's Intersection Cohomology. It doesn't treat perverse sheaves, but it has a good overview of stratified spaces in the first few chapter and develops constructible sheaves and the six functors in Chapter V. You might also try Dimca's Sheaves in Topology, which does talk about perverse sheaves. Neither of these present the most general possible development, but they're pretty accessible.</p> <p>I'll also plug the discussion of stratified spaces in my book-in-progress, though I take a more topological approach that's not particularly oriented toward sheaves. Here's a link: <a href="http://faculty.tcu.edu/gfriedman/IHbook.pdf" rel="nofollow noreferrer">http://faculty.tcu.edu/gfriedman/IHbook.pdf</a></p> Answer 3: <p>Consult the book <em>"Sheaves on manifolds"</em> by Kashiwara and Schapira. It's a hard nut to crack, but it is the most efficient presentation I've seen. </p> <p>In Chapter 8, they work with stratifications whose strata satisfy the <em>$\mu$-regularity condition</em>. This is equivalent to the Verdier regularity condition which in turn is a bit stronger than the Whitney condition b. </p> <p>One can show that any subanalytic set admits stratifications satisfying the $\mu$-regularity conditions. In particular, this condition implies that such stratified spaces admit (many) stratified Morse functions.</p> Answer 4: <p>A bit more reader friendly than Kashiwara-Schapira is Borel's Intersection Cohomology. It doesn't treat perverse sheaves, but it has a good overview of stratified spaces in the first few chapter and develops constructible sheaves and the six functors in Chapter V. You might also try Dimca's Sheaves in Topology, which does talk about perverse sheaves. Neither of these present the most general possible development, but they're pretty accessible.</p> <p>I'll also plug the discussion of stratified spaces in my book-in-progress, though I take a more topological approach that's not particularly oriented toward sheaves. Here's a link: <a href="http://faculty.tcu.edu/gfriedman/IHbook.pdf" rel="nofollow noreferrer">http://faculty.tcu.edu/gfriedman/IHbook.pdf</a></p> Answer 5: <p>Consult the book <em>"Sheaves on manifolds"</em> by Kashiwara and Schapira. It's a hard nut to crack, but it is the most efficient presentation I've seen. </p> <p>In Chapter 8, they work with stratifications whose strata satisfy the <em>$\mu$-regularity condition</em>. This is equivalent to the Verdier regularity condition which in turn is a bit stronger than the Whitney condition b. </p> <p>One can show that any subanalytic set admits stratifications satisfying the $\mu$-regularity conditions. In particular, this condition implies that such stratified spaces admit (many) stratified Morse functions.</p> Answer 6: <p>A bit more reader friendly than Kashiwara-Schapira is Borel's Intersection Cohomology. It doesn't treat perverse sheaves, but it has a good overview of stratified spaces in the first few chapter and develops constructible sheaves and the six functors in Chapter V. You might also try Dimca's Sheaves in Topology, which does talk about perverse sheaves. Neither of these present the most general possible development, but they're pretty accessible.</p> <p>I'll also plug the discussion of stratified spaces in my book-in-progress, though I take a more topological approach that's not particularly oriented toward sheaves. Here's a link: <a href="http://faculty.tcu.edu/gfriedman/IHbook.pdf" rel="nofollow noreferrer">http://faculty.tcu.edu/gfriedman/IHbook.pdf</a></p>
Generalization of winding number to higher dimensions
259,054
14
https://mathoverflow.net/questions/259054/generalization-of-winding-number-to-higher-dimensions
<blockquote> <p>Is there a natural geometric generalization of the <a href="https://en.wikipedia.org/wiki/Winding_number" rel="noreferrer">winding number</a> to higher dimensions?</p> </blockquote> <p>I know it primarily as an important and useful index for closed, plane curves (e.g., the Jordan Curve Theorem), and for its role in Cauchy's theorem integrating holomorphic functions. I would be interested to learn of generalizations that essentially replace the role of the circle $\mathbb{S}^1$ with $\mathbb{S}^n$.</p> <p>I've encountered references to the <a href="https://en.wikipedia.org/wiki/Fredholm_operator" rel="noreferrer">Fredholm index</a>, the <a href="https://www.encyclopediaofmath.org/index.php/Pontryagin_number" rel="noreferrer">Pontryagin index</a>, and to <a href="https://en.wikipedia.org/wiki/Bott_periodicity_theorem" rel="noreferrer">Bott periodicity</a>, but none seem to be straightforward geometric generalizations of winding number.</p> <p>This is an entirely naive question, and references and high-level descriptions would be appreciated, and more than suffice.</p>
[ { "answer_id": 259055, "score": 28, "text": "<p>This is a very naive answer which I am sure you already considered, but isn't the most obvious generalization just given by the topological degree (<a href=\"https://en.wikipedia.org/wiki/Degree_of_a_continuous_mapping\" rel=\"noreferrer\">https://en.wikipedia.org/wiki/Degree_of_a_continuous_mapping</a>)?</p>\n\n<p>The winding number of $f:S^1\\rightarrow \\mathbb{R}^2$ around $p$ is just the degree of the composition of $f$ with the radial projection from $p$, considered as a map from $S^1$ to $S^1$. It is obvious how to do the same thing for general $n$.</p>\n\n<p>(This should surely just be a comment.)</p>\n" }, { "answer_id": 259056, "score": 14, "text": "<p>A smooth function $f$ with image on the unit circle in $\\mathbb{C}$ has <em>winding number:</em>\n$$\\text{wind} f=\\frac1{2\\pi i}\\int f'\\bar{f}=\\frac1{2\\pi i}\\sum\\hat{f'}(n)\\hat{f}(n)=\\sum n\\vert\\hat{f}(n)\\vert^2.$$\nThis formulation allows generalizing the winding number to higher dimensions (eg. $f:\\mathbb{S}^n\\rightarrow\\mathbb{S}^n$) for functions in Sobolev spaces $H^{1/2}$. Notice the natural fit of this space via the characterization of $f\\in H^{1/2}$ through its Fourier coefficients:\n$$\\sum \\vert n\\vert\\,\\vert\\hat{f}(n)\\vert^2&lt;\\infty.$$\nThe first result in this direction was due to L. Boutet de Monvel and O. Gabber. Since then the concept was extended as \"degree of a map\" (obviously the Topologist were aware of <em>topological degree</em>) much further to VMO spaces, etc. A good survey of results and developements on the subject can be found here:</p>\n\n<p>Haïm Brezis, <em>New questions related to the topological degree</em>, The unity of mathematics, 137–154, Progr. Math. 244, Birkhäuser Boston, Boston, MA, 2006</p>\n" }, { "answer_id": 297488, "score": 4, "text": "<p>The linking number is one of a natural generalizations of the winding number, see my answer to a related question: <a href=\"https://mathoverflow.net/a/297440/121665\">https://mathoverflow.net/a/297440/121665</a></p>\n" }, { "answer_id": 297485, "score": 3, "text": "<p>Perhaps it is worth mentioning that in their book \"Mapping Degree Theory\" Outerelo and Ruiz give a general definition of the winding number based on the topological (Brouwer-Kronecker) degree, which was already mentioned in a previous answer. The point being that, while it is a special case of the mapping degree, it is a generalization of the usual winding number that is interesting on its own.</p>\n\n<p><em>Theorem-Definition</em>: Let $U\\subseteq\\mathbb{R}^n$ be bounded and open, let $X:=\\partial U$ be the boundary hypersurface, let $f\\in C^0(X,\\mathbb{R}^n)$, let $\\bar{f}:\\bar{U}\\to\\mathbb{R}^n$ be a continuous extension of $f$ (the set of such extensions is non-empty by Tietze's extension theorem), and let $p\\in\\mathbb{R}^n\\setminus f(X)$. Then $\\deg(\\bar{f},U,a)$ does not depend on the choice of continuous extension $\\bar{f}$, and one defines\n$$\nw(f,a):=\\deg(\\bar{f},U,a)\n$$</p>\n\n<p>The winding number inherits many of the nice properties of the topological degree : being locally constant on the codomain, homotopy invariance, being $0$ for points outside of $\\bar{f}(\\bar{U})$ and others. More importantly, the smooth version of the topological degree can be recovered from this notion of a winding number.</p>\n" }, { "answer_id": 324677, "score": 2, "text": "<p>In your question you mentioned the word \"Fredholm index\".</p>\n\n<p>So I would like to say that in the circle case there are two different interpretations of Fredholm index of certain linear operators in terms of winding number. So it would be interesting to consider a possible generalization of these <span class=\"math-container\">$1$</span> dimensional facts to higher dimensional spheres. </p>\n\n<p>1)If I remember the following theorem correctly, there is a Theorem by Veku in \"Topology and analysis, the Atiyah-Singer index formula and gauge-theoretic physics, by B. Booss and D. D. Bleecker\" which says:</p>\n\n<p><strong>Theorem:</strong> If <span class=\"math-container\">$X$</span> is a non vanishing vector field along <span class=\"math-container\">$S^1\\subset \\mathbb{R}^2$</span>, not necessarily tangent to <span class=\"math-container\">$S^1$</span>, then the fredholm index of the pair operator <span class=\"math-container\">$(\\Delta, \\partial/\\partial X)$</span> is equal to the \"winding number of <span class=\"math-container\">$X$</span>. In the above pair operator <span class=\"math-container\">$\\Delta$</span> is the standard Laplace operator on the interior of circle and the derivational operator <span class=\"math-container\">$\\partial/\\partial X$</span> is restricted to the boundary.</p>\n\n<p>So it would be a good idea to consider an appropriate generalization of this fact. For every smooth self map on sphere one can consider an arbitrary extension to whole <span class=\"math-container\">$\\mathbb{R}^{n+1}$</span> and try to find an appropriate generalization. However an immediate plain generalization is not true but one should consider a modified generalization. I was thinking to this question about 7 years ago and I observed that a plain generalization is not true since the corresponding operator on <span class=\"math-container\">$S^3$</span> is not a Fredholm operator. I had intension to discuss these materials in my following talk I presented in Timisoara but I did not had enough time to present all of the materials of this abstract, since my time was 20 minutes:</p>\n\n<p><a href=\"http://at.yorku.ca/c/b/d/z/37.htm\" rel=\"nofollow noreferrer\">http://at.yorku.ca/c/b/d/z/37.htm</a></p>\n\n<p>(Sorry, the talk abstract I wrote is very snafu)</p>\n\n<p>2) Many years ago I learned from a speaker in Non commutative geometry<br>\nthat the multiplication by a non vanishing complex map <span class=\"math-container\">$f$</span> on <span class=\"math-container\">$S^1$</span> is a fredholm operator on <span class=\"math-container\">$L^1(S^1)$</span> whose index is <span class=\"math-container\">$- W(f)$</span>, the winding number of <span class=\"math-container\">$f$</span>. So it would be interesting to consider an <span class=\"math-container\">$S^3$</span> analogy or even more a compact Lie group analogy.</p>\n\n<p>The following paper contains an explanation of a similar situation</p>\n\n<p><a href=\"http://www.ams.org/journals/proc/2005-133-05/S0002-9939-04-07642-7/S0002-9939-04-07642-7.pdf\" rel=\"nofollow noreferrer\">http://www.ams.org/journals/proc/2005-133-05/S0002-9939-04-07642-7/S0002-9939-04-07642-7.pdf</a></p>\n" }, { "answer_id": 259055, "score": 28, "text": "<p>This is a very naive answer which I am sure you already considered, but isn't the most obvious generalization just given by the topological degree (<a href=\"https://en.wikipedia.org/wiki/Degree_of_a_continuous_mapping\" rel=\"noreferrer\">https://en.wikipedia.org/wiki/Degree_of_a_continuous_mapping</a>)?</p>\n\n<p>The winding number of $f:S^1\\rightarrow \\mathbb{R}^2$ around $p$ is just the degree of the composition of $f$ with the radial projection from $p$, considered as a map from $S^1$ to $S^1$. It is obvious how to do the same thing for general $n$.</p>\n\n<p>(This should surely just be a comment.)</p>\n" }, { "answer_id": 259056, "score": 14, "text": "<p>A smooth function $f$ with image on the unit circle in $\\mathbb{C}$ has <em>winding number:</em>\n$$\\text{wind} f=\\frac1{2\\pi i}\\int f'\\bar{f}=\\frac1{2\\pi i}\\sum\\hat{f'}(n)\\hat{f}(n)=\\sum n\\vert\\hat{f}(n)\\vert^2.$$\nThis formulation allows generalizing the winding number to higher dimensions (eg. $f:\\mathbb{S}^n\\rightarrow\\mathbb{S}^n$) for functions in Sobolev spaces $H^{1/2}$. Notice the natural fit of this space via the characterization of $f\\in H^{1/2}$ through its Fourier coefficients:\n$$\\sum \\vert n\\vert\\,\\vert\\hat{f}(n)\\vert^2&lt;\\infty.$$\nThe first result in this direction was due to L. Boutet de Monvel and O. Gabber. Since then the concept was extended as \"degree of a map\" (obviously the Topologist were aware of <em>topological degree</em>) much further to VMO spaces, etc. A good survey of results and developements on the subject can be found here:</p>\n\n<p>Haïm Brezis, <em>New questions related to the topological degree</em>, The unity of mathematics, 137–154, Progr. Math. 244, Birkhäuser Boston, Boston, MA, 2006</p>\n" }, { "answer_id": 297488, "score": 4, "text": "<p>The linking number is one of a natural generalizations of the winding number, see my answer to a related question: <a href=\"https://mathoverflow.net/a/297440/121665\">https://mathoverflow.net/a/297440/121665</a></p>\n" }, { "answer_id": 297485, "score": 3, "text": "<p>Perhaps it is worth mentioning that in their book \"Mapping Degree Theory\" Outerelo and Ruiz give a general definition of the winding number based on the topological (Brouwer-Kronecker) degree, which was already mentioned in a previous answer. The point being that, while it is a special case of the mapping degree, it is a generalization of the usual winding number that is interesting on its own.</p>\n\n<p><em>Theorem-Definition</em>: Let $U\\subseteq\\mathbb{R}^n$ be bounded and open, let $X:=\\partial U$ be the boundary hypersurface, let $f\\in C^0(X,\\mathbb{R}^n)$, let $\\bar{f}:\\bar{U}\\to\\mathbb{R}^n$ be a continuous extension of $f$ (the set of such extensions is non-empty by Tietze's extension theorem), and let $p\\in\\mathbb{R}^n\\setminus f(X)$. Then $\\deg(\\bar{f},U,a)$ does not depend on the choice of continuous extension $\\bar{f}$, and one defines\n$$\nw(f,a):=\\deg(\\bar{f},U,a)\n$$</p>\n\n<p>The winding number inherits many of the nice properties of the topological degree : being locally constant on the codomain, homotopy invariance, being $0$ for points outside of $\\bar{f}(\\bar{U})$ and others. More importantly, the smooth version of the topological degree can be recovered from this notion of a winding number.</p>\n" }, { "answer_id": 324677, "score": 2, "text": "<p>In your question you mentioned the word \"Fredholm index\".</p>\n\n<p>So I would like to say that in the circle case there are two different interpretations of Fredholm index of certain linear operators in terms of winding number. So it would be interesting to consider a possible generalization of these <span class=\"math-container\">$1$</span> dimensional facts to higher dimensional spheres. </p>\n\n<p>1)If I remember the following theorem correctly, there is a Theorem by Veku in \"Topology and analysis, the Atiyah-Singer index formula and gauge-theoretic physics, by B. Booss and D. D. Bleecker\" which says:</p>\n\n<p><strong>Theorem:</strong> If <span class=\"math-container\">$X$</span> is a non vanishing vector field along <span class=\"math-container\">$S^1\\subset \\mathbb{R}^2$</span>, not necessarily tangent to <span class=\"math-container\">$S^1$</span>, then the fredholm index of the pair operator <span class=\"math-container\">$(\\Delta, \\partial/\\partial X)$</span> is equal to the \"winding number of <span class=\"math-container\">$X$</span>. In the above pair operator <span class=\"math-container\">$\\Delta$</span> is the standard Laplace operator on the interior of circle and the derivational operator <span class=\"math-container\">$\\partial/\\partial X$</span> is restricted to the boundary.</p>\n\n<p>So it would be a good idea to consider an appropriate generalization of this fact. For every smooth self map on sphere one can consider an arbitrary extension to whole <span class=\"math-container\">$\\mathbb{R}^{n+1}$</span> and try to find an appropriate generalization. However an immediate plain generalization is not true but one should consider a modified generalization. I was thinking to this question about 7 years ago and I observed that a plain generalization is not true since the corresponding operator on <span class=\"math-container\">$S^3$</span> is not a Fredholm operator. I had intension to discuss these materials in my following talk I presented in Timisoara but I did not had enough time to present all of the materials of this abstract, since my time was 20 minutes:</p>\n\n<p><a href=\"http://at.yorku.ca/c/b/d/z/37.htm\" rel=\"nofollow noreferrer\">http://at.yorku.ca/c/b/d/z/37.htm</a></p>\n\n<p>(Sorry, the talk abstract I wrote is very snafu)</p>\n\n<p>2) Many years ago I learned from a speaker in Non commutative geometry<br>\nthat the multiplication by a non vanishing complex map <span class=\"math-container\">$f$</span> on <span class=\"math-container\">$S^1$</span> is a fredholm operator on <span class=\"math-container\">$L^1(S^1)$</span> whose index is <span class=\"math-container\">$- W(f)$</span>, the winding number of <span class=\"math-container\">$f$</span>. So it would be interesting to consider an <span class=\"math-container\">$S^3$</span> analogy or even more a compact Lie group analogy.</p>\n\n<p>The following paper contains an explanation of a similar situation</p>\n\n<p><a href=\"http://www.ams.org/journals/proc/2005-133-05/S0002-9939-04-07642-7/S0002-9939-04-07642-7.pdf\" rel=\"nofollow noreferrer\">http://www.ams.org/journals/proc/2005-133-05/S0002-9939-04-07642-7/S0002-9939-04-07642-7.pdf</a></p>\n" }, { "answer_id": 259055, "score": 28, "text": "<p>This is a very naive answer which I am sure you already considered, but isn't the most obvious generalization just given by the topological degree (<a href=\"https://en.wikipedia.org/wiki/Degree_of_a_continuous_mapping\" rel=\"noreferrer\">https://en.wikipedia.org/wiki/Degree_of_a_continuous_mapping</a>)?</p>\n\n<p>The winding number of $f:S^1\\rightarrow \\mathbb{R}^2$ around $p$ is just the degree of the composition of $f$ with the radial projection from $p$, considered as a map from $S^1$ to $S^1$. It is obvious how to do the same thing for general $n$.</p>\n\n<p>(This should surely just be a comment.)</p>\n" }, { "answer_id": 259056, "score": 14, "text": "<p>A smooth function $f$ with image on the unit circle in $\\mathbb{C}$ has <em>winding number:</em>\n$$\\text{wind} f=\\frac1{2\\pi i}\\int f'\\bar{f}=\\frac1{2\\pi i}\\sum\\hat{f'}(n)\\hat{f}(n)=\\sum n\\vert\\hat{f}(n)\\vert^2.$$\nThis formulation allows generalizing the winding number to higher dimensions (eg. $f:\\mathbb{S}^n\\rightarrow\\mathbb{S}^n$) for functions in Sobolev spaces $H^{1/2}$. Notice the natural fit of this space via the characterization of $f\\in H^{1/2}$ through its Fourier coefficients:\n$$\\sum \\vert n\\vert\\,\\vert\\hat{f}(n)\\vert^2&lt;\\infty.$$\nThe first result in this direction was due to L. Boutet de Monvel and O. Gabber. Since then the concept was extended as \"degree of a map\" (obviously the Topologist were aware of <em>topological degree</em>) much further to VMO spaces, etc. A good survey of results and developements on the subject can be found here:</p>\n\n<p>Haïm Brezis, <em>New questions related to the topological degree</em>, The unity of mathematics, 137–154, Progr. Math. 244, Birkhäuser Boston, Boston, MA, 2006</p>\n" }, { "answer_id": 297488, "score": 4, "text": "<p>The linking number is one of a natural generalizations of the winding number, see my answer to a related question: <a href=\"https://mathoverflow.net/a/297440/121665\">https://mathoverflow.net/a/297440/121665</a></p>\n" }, { "answer_id": 297485, "score": 3, "text": "<p>Perhaps it is worth mentioning that in their book \"Mapping Degree Theory\" Outerelo and Ruiz give a general definition of the winding number based on the topological (Brouwer-Kronecker) degree, which was already mentioned in a previous answer. The point being that, while it is a special case of the mapping degree, it is a generalization of the usual winding number that is interesting on its own.</p>\n\n<p><em>Theorem-Definition</em>: Let $U\\subseteq\\mathbb{R}^n$ be bounded and open, let $X:=\\partial U$ be the boundary hypersurface, let $f\\in C^0(X,\\mathbb{R}^n)$, let $\\bar{f}:\\bar{U}\\to\\mathbb{R}^n$ be a continuous extension of $f$ (the set of such extensions is non-empty by Tietze's extension theorem), and let $p\\in\\mathbb{R}^n\\setminus f(X)$. Then $\\deg(\\bar{f},U,a)$ does not depend on the choice of continuous extension $\\bar{f}$, and one defines\n$$\nw(f,a):=\\deg(\\bar{f},U,a)\n$$</p>\n\n<p>The winding number inherits many of the nice properties of the topological degree : being locally constant on the codomain, homotopy invariance, being $0$ for points outside of $\\bar{f}(\\bar{U})$ and others. More importantly, the smooth version of the topological degree can be recovered from this notion of a winding number.</p>\n" }, { "answer_id": 324677, "score": 2, "text": "<p>In your question you mentioned the word \"Fredholm index\".</p>\n\n<p>So I would like to say that in the circle case there are two different interpretations of Fredholm index of certain linear operators in terms of winding number. So it would be interesting to consider a possible generalization of these <span class=\"math-container\">$1$</span> dimensional facts to higher dimensional spheres. </p>\n\n<p>1)If I remember the following theorem correctly, there is a Theorem by Veku in \"Topology and analysis, the Atiyah-Singer index formula and gauge-theoretic physics, by B. Booss and D. D. Bleecker\" which says:</p>\n\n<p><strong>Theorem:</strong> If <span class=\"math-container\">$X$</span> is a non vanishing vector field along <span class=\"math-container\">$S^1\\subset \\mathbb{R}^2$</span>, not necessarily tangent to <span class=\"math-container\">$S^1$</span>, then the fredholm index of the pair operator <span class=\"math-container\">$(\\Delta, \\partial/\\partial X)$</span> is equal to the \"winding number of <span class=\"math-container\">$X$</span>. In the above pair operator <span class=\"math-container\">$\\Delta$</span> is the standard Laplace operator on the interior of circle and the derivational operator <span class=\"math-container\">$\\partial/\\partial X$</span> is restricted to the boundary.</p>\n\n<p>So it would be a good idea to consider an appropriate generalization of this fact. For every smooth self map on sphere one can consider an arbitrary extension to whole <span class=\"math-container\">$\\mathbb{R}^{n+1}$</span> and try to find an appropriate generalization. However an immediate plain generalization is not true but one should consider a modified generalization. I was thinking to this question about 7 years ago and I observed that a plain generalization is not true since the corresponding operator on <span class=\"math-container\">$S^3$</span> is not a Fredholm operator. I had intension to discuss these materials in my following talk I presented in Timisoara but I did not had enough time to present all of the materials of this abstract, since my time was 20 minutes:</p>\n\n<p><a href=\"http://at.yorku.ca/c/b/d/z/37.htm\" rel=\"nofollow noreferrer\">http://at.yorku.ca/c/b/d/z/37.htm</a></p>\n\n<p>(Sorry, the talk abstract I wrote is very snafu)</p>\n\n<p>2) Many years ago I learned from a speaker in Non commutative geometry<br>\nthat the multiplication by a non vanishing complex map <span class=\"math-container\">$f$</span> on <span class=\"math-container\">$S^1$</span> is a fredholm operator on <span class=\"math-container\">$L^1(S^1)$</span> whose index is <span class=\"math-container\">$- W(f)$</span>, the winding number of <span class=\"math-container\">$f$</span>. So it would be interesting to consider an <span class=\"math-container\">$S^3$</span> analogy or even more a compact Lie group analogy.</p>\n\n<p>The following paper contains an explanation of a similar situation</p>\n\n<p><a href=\"http://www.ams.org/journals/proc/2005-133-05/S0002-9939-04-07642-7/S0002-9939-04-07642-7.pdf\" rel=\"nofollow noreferrer\">http://www.ams.org/journals/proc/2005-133-05/S0002-9939-04-07642-7/S0002-9939-04-07642-7.pdf</a></p>\n" } ]
Question: Generalization of winding number to higher dimensions Answer 1: <p>This is a very naive answer which I am sure you already considered, but isn't the most obvious generalization just given by the topological degree (<a href="https://en.wikipedia.org/wiki/Degree_of_a_continuous_mapping" rel="noreferrer">https://en.wikipedia.org/wiki/Degree_of_a_continuous_mapping</a>)?</p> <p>The winding number of $f:S^1\rightarrow \mathbb{R}^2$ around $p$ is just the degree of the composition of $f$ with the radial projection from $p$, considered as a map from $S^1$ to $S^1$. It is obvious how to do the same thing for general $n$.</p> <p>(This should surely just be a comment.)</p> Answer 2: <p>A smooth function $f$ with image on the unit circle in $\mathbb{C}$ has <em>winding number:</em> $$\text{wind} f=\frac1{2\pi i}\int f'\bar{f}=\frac1{2\pi i}\sum\hat{f'}(n)\hat{f}(n)=\sum n\vert\hat{f}(n)\vert^2.$$ This formulation allows generalizing the winding number to higher dimensions (eg. $f:\mathbb{S}^n\rightarrow\mathbb{S}^n$) for functions in Sobolev spaces $H^{1/2}$. Notice the natural fit of this space via the characterization of $f\in H^{1/2}$ through its Fourier coefficients: $$\sum \vert n\vert\,\vert\hat{f}(n)\vert^2&lt;\infty.$$ The first result in this direction was due to L. Boutet de Monvel and O. Gabber. Since then the concept was extended as "degree of a map" (obviously the Topologist were aware of <em>topological degree</em>) much further to VMO spaces, etc. A good survey of results and developements on the subject can be found here:</p> <p>Haïm Brezis, <em>New questions related to the topological degree</em>, The unity of mathematics, 137–154, Progr. Math. 244, Birkhäuser Boston, Boston, MA, 2006</p> Answer 3: <p>The linking number is one of a natural generalizations of the winding number, see my answer to a related question: <a href="https://mathoverflow.net/a/297440/121665">https://mathoverflow.net/a/297440/121665</a></p> Answer 4: <p>Perhaps it is worth mentioning that in their book "Mapping Degree Theory" Outerelo and Ruiz give a general definition of the winding number based on the topological (Brouwer-Kronecker) degree, which was already mentioned in a previous answer. The point being that, while it is a special case of the mapping degree, it is a generalization of the usual winding number that is interesting on its own.</p> <p><em>Theorem-Definition</em>: Let $U\subseteq\mathbb{R}^n$ be bounded and open, let $X:=\partial U$ be the boundary hypersurface, let $f\in C^0(X,\mathbb{R}^n)$, let $\bar{f}:\bar{U}\to\mathbb{R}^n$ be a continuous extension of $f$ (the set of such extensions is non-empty by Tietze's extension theorem), and let $p\in\mathbb{R}^n\setminus f(X)$. Then $\deg(\bar{f},U,a)$ does not depend on the choice of continuous extension $\bar{f}$, and one defines $$ w(f,a):=\deg(\bar{f},U,a) $$</p> <p>The winding number inherits many of the nice properties of the topological degree : being locally constant on the codomain, homotopy invariance, being $0$ for points outside of $\bar{f}(\bar{U})$ and others. More importantly, the smooth version of the topological degree can be recovered from this notion of a winding number.</p> Answer 5: <p>In your question you mentioned the word "Fredholm index".</p> <p>So I would like to say that in the circle case there are two different interpretations of Fredholm index of certain linear operators in terms of winding number. So it would be interesting to consider a possible generalization of these <span class="math-container">$1$</span> dimensional facts to higher dimensional spheres. </p> <p>1)If I remember the following theorem correctly, there is a Theorem by Veku in "Topology and analysis, the Atiyah-Singer index formula and gauge-theoretic physics, by B. Booss and D. D. Bleecker" which says:</p> <p><strong>Theorem:</strong> If <span class="math-container">$X$</span> is a non vanishing vector field along <span class="math-container">$S^1\subset \mathbb{R}^2$</span>, not necessarily tangent to <span class="math-container">$S^1$</span>, then the fredholm index of the pair operator <span class="math-container">$(\Delta, \partial/\partial X)$</span> is equal to the "winding number of <span class="math-container">$X$</span>. In the above pair operator <span class="math-container">$\Delta$</span> is the standard Laplace operator on the interior of circle and the derivational operator <span class="math-container">$\partial/\partial X$</span> is restricted to the boundary.</p> <p>So it would be a good idea to consider an appropriate generalization of this fact. For every smooth self map on sphere one can consider an arbitrary extension to whole <span class="math-container">$\mathbb{R}^{n+1}$</span> and try to find an appropriate generalization. However an immediate plain generalization is not true but one should consider a modified generalization. I was thinking to this question about 7 years ago and I observed that a plain generalization is not true since the corresponding operator on <span class="math-container">$S^3$</span> is not a Fredholm operator. I had intension to discuss these materials in my following talk I presented in Timisoara but I did not had enough time to present all of the materials of this abstract, since my time was 20 minutes:</p> <p><a href="http://at.yorku.ca/c/b/d/z/37.htm" rel="nofollow noreferrer">http://at.yorku.ca/c/b/d/z/37.htm</a></p> <p>(Sorry, the talk abstract I wrote is very snafu)</p> <p>2) Many years ago I learned from a speaker in Non commutative geometry<br> that the multiplication by a non vanishing complex map <span class="math-container">$f$</span> on <span class="math-container">$S^1$</span> is a fredholm operator on <span class="math-container">$L^1(S^1)$</span> whose index is <span class="math-container">$- W(f)$</span>, the winding number of <span class="math-container">$f$</span>. So it would be interesting to consider an <span class="math-container">$S^3$</span> analogy or even more a compact Lie group analogy.</p> <p>The following paper contains an explanation of a similar situation</p> <p><a href="http://www.ams.org/journals/proc/2005-133-05/S0002-9939-04-07642-7/S0002-9939-04-07642-7.pdf" rel="nofollow noreferrer">http://www.ams.org/journals/proc/2005-133-05/S0002-9939-04-07642-7/S0002-9939-04-07642-7.pdf</a></p> Answer 6: <p>This is a very naive answer which I am sure you already considered, but isn't the most obvious generalization just given by the topological degree (<a href="https://en.wikipedia.org/wiki/Degree_of_a_continuous_mapping" rel="noreferrer">https://en.wikipedia.org/wiki/Degree_of_a_continuous_mapping</a>)?</p> <p>The winding number of $f:S^1\rightarrow \mathbb{R}^2$ around $p$ is just the degree of the composition of $f$ with the radial projection from $p$, considered as a map from $S^1$ to $S^1$. It is obvious how to do the same thing for general $n$.</p> <p>(This should surely just be a comment.)</p> Answer 7: <p>A smooth function $f$ with image on the unit circle in $\mathbb{C}$ has <em>winding number:</em> $$\text{wind} f=\frac1{2\pi i}\int f'\bar{f}=\frac1{2\pi i}\sum\hat{f'}(n)\hat{f}(n)=\sum n\vert\hat{f}(n)\vert^2.$$ This formulation allows generalizing the winding number to higher dimensions (eg. $f:\mathbb{S}^n\rightarrow\mathbb{S}^n$) for functions in Sobolev spaces $H^{1/2}$. Notice the natural fit of this space via the characterization of $f\in H^{1/2}$ through its Fourier coefficients: $$\sum \vert n\vert\,\vert\hat{f}(n)\vert^2&lt;\infty.$$ The first result in this direction was due to L. Boutet de Monvel and O. Gabber. Since then the concept was extended as "degree of a map" (obviously the Topologist were aware of <em>topological degree</em>) much further to VMO spaces, etc. A good survey of results and developements on the subject can be found here:</p> <p>Haïm Brezis, <em>New questions related to the topological degree</em>, The unity of mathematics, 137–154, Progr. Math. 244, Birkhäuser Boston, Boston, MA, 2006</p> Answer 8: <p>The linking number is one of a natural generalizations of the winding number, see my answer to a related question: <a href="https://mathoverflow.net/a/297440/121665">https://mathoverflow.net/a/297440/121665</a></p> Answer 9: <p>Perhaps it is worth mentioning that in their book "Mapping Degree Theory" Outerelo and Ruiz give a general definition of the winding number based on the topological (Brouwer-Kronecker) degree, which was already mentioned in a previous answer. The point being that, while it is a special case of the mapping degree, it is a generalization of the usual winding number that is interesting on its own.</p> <p><em>Theorem-Definition</em>: Let $U\subseteq\mathbb{R}^n$ be bounded and open, let $X:=\partial U$ be the boundary hypersurface, let $f\in C^0(X,\mathbb{R}^n)$, let $\bar{f}:\bar{U}\to\mathbb{R}^n$ be a continuous extension of $f$ (the set of such extensions is non-empty by Tietze's extension theorem), and let $p\in\mathbb{R}^n\setminus f(X)$. Then $\deg(\bar{f},U,a)$ does not depend on the choice of continuous extension $\bar{f}$, and one defines $$ w(f,a):=\deg(\bar{f},U,a) $$</p> <p>The winding number inherits many of the nice properties of the topological degree : being locally constant on the codomain, homotopy invariance, being $0$ for points outside of $\bar{f}(\bar{U})$ and others. More importantly, the smooth version of the topological degree can be recovered from this notion of a winding number.</p> Answer 10: <p>In your question you mentioned the word "Fredholm index".</p> <p>So I would like to say that in the circle case there are two different interpretations of Fredholm index of certain linear operators in terms of winding number. So it would be interesting to consider a possible generalization of these <span class="math-container">$1$</span> dimensional facts to higher dimensional spheres. </p> <p>1)If I remember the following theorem correctly, there is a Theorem by Veku in "Topology and analysis, the Atiyah-Singer index formula and gauge-theoretic physics, by B. Booss and D. D. Bleecker" which says:</p> <p><strong>Theorem:</strong> If <span class="math-container">$X$</span> is a non vanishing vector field along <span class="math-container">$S^1\subset \mathbb{R}^2$</span>, not necessarily tangent to <span class="math-container">$S^1$</span>, then the fredholm index of the pair operator <span class="math-container">$(\Delta, \partial/\partial X)$</span> is equal to the "winding number of <span class="math-container">$X$</span>. In the above pair operator <span class="math-container">$\Delta$</span> is the standard Laplace operator on the interior of circle and the derivational operator <span class="math-container">$\partial/\partial X$</span> is restricted to the boundary.</p> <p>So it would be a good idea to consider an appropriate generalization of this fact. For every smooth self map on sphere one can consider an arbitrary extension to whole <span class="math-container">$\mathbb{R}^{n+1}$</span> and try to find an appropriate generalization. However an immediate plain generalization is not true but one should consider a modified generalization. I was thinking to this question about 7 years ago and I observed that a plain generalization is not true since the corresponding operator on <span class="math-container">$S^3$</span> is not a Fredholm operator. I had intension to discuss these materials in my following talk I presented in Timisoara but I did not had enough time to present all of the materials of this abstract, since my time was 20 minutes:</p> <p><a href="http://at.yorku.ca/c/b/d/z/37.htm" rel="nofollow noreferrer">http://at.yorku.ca/c/b/d/z/37.htm</a></p> <p>(Sorry, the talk abstract I wrote is very snafu)</p> <p>2) Many years ago I learned from a speaker in Non commutative geometry<br> that the multiplication by a non vanishing complex map <span class="math-container">$f$</span> on <span class="math-container">$S^1$</span> is a fredholm operator on <span class="math-container">$L^1(S^1)$</span> whose index is <span class="math-container">$- W(f)$</span>, the winding number of <span class="math-container">$f$</span>. So it would be interesting to consider an <span class="math-container">$S^3$</span> analogy or even more a compact Lie group analogy.</p> <p>The following paper contains an explanation of a similar situation</p> <p><a href="http://www.ams.org/journals/proc/2005-133-05/S0002-9939-04-07642-7/S0002-9939-04-07642-7.pdf" rel="nofollow noreferrer">http://www.ams.org/journals/proc/2005-133-05/S0002-9939-04-07642-7/S0002-9939-04-07642-7.pdf</a></p> Answer 11: <p>This is a very naive answer which I am sure you already considered, but isn't the most obvious generalization just given by the topological degree (<a href="https://en.wikipedia.org/wiki/Degree_of_a_continuous_mapping" rel="noreferrer">https://en.wikipedia.org/wiki/Degree_of_a_continuous_mapping</a>)?</p> <p>The winding number of $f:S^1\rightarrow \mathbb{R}^2$ around $p$ is just the degree of the composition of $f$ with the radial projection from $p$, considered as a map from $S^1$ to $S^1$. It is obvious how to do the same thing for general $n$.</p> <p>(This should surely just be a comment.)</p> Answer 12: <p>A smooth function $f$ with image on the unit circle in $\mathbb{C}$ has <em>winding number:</em> $$\text{wind} f=\frac1{2\pi i}\int f'\bar{f}=\frac1{2\pi i}\sum\hat{f'}(n)\hat{f}(n)=\sum n\vert\hat{f}(n)\vert^2.$$ This formulation allows generalizing the winding number to higher dimensions (eg. $f:\mathbb{S}^n\rightarrow\mathbb{S}^n$) for functions in Sobolev spaces $H^{1/2}$. Notice the natural fit of this space via the characterization of $f\in H^{1/2}$ through its Fourier coefficients: $$\sum \vert n\vert\,\vert\hat{f}(n)\vert^2&lt;\infty.$$ The first result in this direction was due to L. Boutet de Monvel and O. Gabber. Since then the concept was extended as "degree of a map" (obviously the Topologist were aware of <em>topological degree</em>) much further to VMO spaces, etc. A good survey of results and developements on the subject can be found here:</p> <p>Haïm Brezis, <em>New questions related to the topological degree</em>, The unity of mathematics, 137–154, Progr. Math. 244, Birkhäuser Boston, Boston, MA, 2006</p> Answer 13: <p>The linking number is one of a natural generalizations of the winding number, see my answer to a related question: <a href="https://mathoverflow.net/a/297440/121665">https://mathoverflow.net/a/297440/121665</a></p> Answer 14: <p>Perhaps it is worth mentioning that in their book "Mapping Degree Theory" Outerelo and Ruiz give a general definition of the winding number based on the topological (Brouwer-Kronecker) degree, which was already mentioned in a previous answer. The point being that, while it is a special case of the mapping degree, it is a generalization of the usual winding number that is interesting on its own.</p> <p><em>Theorem-Definition</em>: Let $U\subseteq\mathbb{R}^n$ be bounded and open, let $X:=\partial U$ be the boundary hypersurface, let $f\in C^0(X,\mathbb{R}^n)$, let $\bar{f}:\bar{U}\to\mathbb{R}^n$ be a continuous extension of $f$ (the set of such extensions is non-empty by Tietze's extension theorem), and let $p\in\mathbb{R}^n\setminus f(X)$. Then $\deg(\bar{f},U,a)$ does not depend on the choice of continuous extension $\bar{f}$, and one defines $$ w(f,a):=\deg(\bar{f},U,a) $$</p> <p>The winding number inherits many of the nice properties of the topological degree : being locally constant on the codomain, homotopy invariance, being $0$ for points outside of $\bar{f}(\bar{U})$ and others. More importantly, the smooth version of the topological degree can be recovered from this notion of a winding number.</p> Answer 15: <p>In your question you mentioned the word "Fredholm index".</p> <p>So I would like to say that in the circle case there are two different interpretations of Fredholm index of certain linear operators in terms of winding number. So it would be interesting to consider a possible generalization of these <span class="math-container">$1$</span> dimensional facts to higher dimensional spheres. </p> <p>1)If I remember the following theorem correctly, there is a Theorem by Veku in "Topology and analysis, the Atiyah-Singer index formula and gauge-theoretic physics, by B. Booss and D. D. Bleecker" which says:</p> <p><strong>Theorem:</strong> If <span class="math-container">$X$</span> is a non vanishing vector field along <span class="math-container">$S^1\subset \mathbb{R}^2$</span>, not necessarily tangent to <span class="math-container">$S^1$</span>, then the fredholm index of the pair operator <span class="math-container">$(\Delta, \partial/\partial X)$</span> is equal to the "winding number of <span class="math-container">$X$</span>. In the above pair operator <span class="math-container">$\Delta$</span> is the standard Laplace operator on the interior of circle and the derivational operator <span class="math-container">$\partial/\partial X$</span> is restricted to the boundary.</p> <p>So it would be a good idea to consider an appropriate generalization of this fact. For every smooth self map on sphere one can consider an arbitrary extension to whole <span class="math-container">$\mathbb{R}^{n+1}$</span> and try to find an appropriate generalization. However an immediate plain generalization is not true but one should consider a modified generalization. I was thinking to this question about 7 years ago and I observed that a plain generalization is not true since the corresponding operator on <span class="math-container">$S^3$</span> is not a Fredholm operator. I had intension to discuss these materials in my following talk I presented in Timisoara but I did not had enough time to present all of the materials of this abstract, since my time was 20 minutes:</p> <p><a href="http://at.yorku.ca/c/b/d/z/37.htm" rel="nofollow noreferrer">http://at.yorku.ca/c/b/d/z/37.htm</a></p> <p>(Sorry, the talk abstract I wrote is very snafu)</p> <p>2) Many years ago I learned from a speaker in Non commutative geometry<br> that the multiplication by a non vanishing complex map <span class="math-container">$f$</span> on <span class="math-container">$S^1$</span> is a fredholm operator on <span class="math-container">$L^1(S^1)$</span> whose index is <span class="math-container">$- W(f)$</span>, the winding number of <span class="math-container">$f$</span>. So it would be interesting to consider an <span class="math-container">$S^3$</span> analogy or even more a compact Lie group analogy.</p> <p>The following paper contains an explanation of a similar situation</p> <p><a href="http://www.ams.org/journals/proc/2005-133-05/S0002-9939-04-07642-7/S0002-9939-04-07642-7.pdf" rel="nofollow noreferrer">http://www.ams.org/journals/proc/2005-133-05/S0002-9939-04-07642-7/S0002-9939-04-07642-7.pdf</a></p>
Conjugate points on cut locus
260,032
14
https://mathoverflow.net/questions/260032/conjugate-points-on-cut-locus
<p>Let $M$ be a Riemannian with nonempty boundary $\partial M$. Define <em>multiplicity</em> of $x\in M$ as the number of minimizing geodesics from $x$ to $\partial M$. </p> <p>The following fact seems to be standard:</p> <blockquote> <p>The set of points with multiplicity $\ge 2$ is dense in the cut locus of $M$ with respect to $\partial M$.</p> </blockquote> <p>Is it proved somewhere?</p> <p><strong>Comments</strong></p> <ul> <li><p>I know how to prove it, but if you have a one-line proof I am very interested.</p></li> <li><p>Any point of multiplicity 1 in the cut locus is conjugate, but not the other way around. Say for a hemisphere, the cut locus contains a single point — it is a conjugate point of multiplicity $\infty$.</p></li> </ul>
[ { "answer_id": 260949, "score": 6, "text": "<p>Intuitively, this seems plausible if $M$ is complete. Let $d:\\partial M \\to [0,\\infty]$ be the distance function to the cut locus. Then $d$ is continuous (by completeness), and $graph(d) \\subset \\partial M\\times [0,\\infty]$ is a copy of $\\partial M$. There is an open subset $V\\subset \\partial M$ on which $d$ is finite. </p>\n\n<p>Then $M$ is obtained by taking $U=\\{ (x,t)\\subset \\partial M\\times [0,\\infty) | t\\leq d(x)\\}$ and taking the quotient of $graph(d_{|V})$ by the identification map to the cut locus.</p>\n\n<p>If the multiplicity $\\geq 2$ points are not dense in the cut locus, then there is an open subset in the cut locus of multiplicity one. This should pull back to an open subset of $graph(d_{|V})$, which maps homeomorophically under the identification map. But an open ball subset of $graph(d_{|V})$ mapping to the cut locus with multiplicity one will have a collar neighborhood homeomorphic to a half-space, which maps homeomorphically to $M$. However, the boundary of the half-space maps to the cut locus, not to $\\partial M$, which is a contradiction to the fact that $M$ is a manifold near the cut locus. </p>\n" }, { "answer_id": 261234, "score": 5, "text": "<p><em>Let me sketch another proof. It use analysis instead of topology. Namely we use existence of flow for continuous vector fields (uniqueness requires more, but we will not use it).</em></p>\n\n<p>Denote by <span class=\"math-container\">$K$</span> the subset in the interior of <span class=\"math-container\">$M$</span> with multiplicity 1\nand let <span class=\"math-container\">$U\\subset K$</span> be its interior. We need to show that <span class=\"math-container\">$U$</span> does not intersect the cut locus.</p>\n\n<p>Consider the function <span class=\"math-container\">$f= \\mathrm{dist}_{∂M}$</span>.\nNote that <span class=\"math-container\">$K$</span> is the set of points where <span class=\"math-container\">$f$</span> is differentiable, <span class=\"math-container\">$|\\nabla_x\\,f|=1$</span> for any <span class=\"math-container\">$x\\in K$</span> and <span class=\"math-container\">$\\nabla f$</span> is a continuous vector field in <span class=\"math-container\">$U$</span>. \nIn particular <span class=\"math-container\">$f$</span>-gradient flow exist in <span class=\"math-container\">$U$</span> (in fact, it is unique, but we do not need it).</p>\n\n<p>Fix <span class=\"math-container\">$x \\in U$</span>.\nMoving <span class=\"math-container\">$x$</span> along the <span class=\"math-container\">$f$</span>-gradient flow, we get a curve <span class=\"math-container\">$\\gamma$</span> such that <span class=\"math-container\">$(f\\circ\\gamma)'\\equiv 1$</span>.\nIn other words, <span class=\"math-container\">$\\gamma$</span> is an extension of the minimizing geodesic from <span class=\"math-container\">$∂M$</span> to <span class=\"math-container\">$x$</span>.\nHence <span class=\"math-container\">$x$</span> does not lie on the cut locus.</p>\n\n<p>Since <span class=\"math-container\">$x\\in U$</span> is arbitrary, the statement follows.</p>\n" }, { "answer_id": 261914, "score": 5, "text": "<p><em>I've got a letter from Stephanie Alexander with a complete answer. Let me summarize it here.</em></p>\n\n<p>The first proof is given in 4.8 of \"Schnittort und konvexe Mengen...\" by Hermann Karcher (1968). (The formulation is slightly weaker, but from the proof proves our statement follows; the idea is the same as in the answer of Ian Agol.)</p>\n\n<p>An other proof was given in \"Decomposition of cut loci\" by Richard Bishop (1977).</p>\n\n<p>Latter it appears as Lemma 2 in \"Distance function and cut loci...\" by Franz-Erich Wolter (1979). The proof is basically\nthe same as Karcher's.</p>\n" }, { "answer_id": 260949, "score": 6, "text": "<p>Intuitively, this seems plausible if $M$ is complete. Let $d:\\partial M \\to [0,\\infty]$ be the distance function to the cut locus. Then $d$ is continuous (by completeness), and $graph(d) \\subset \\partial M\\times [0,\\infty]$ is a copy of $\\partial M$. There is an open subset $V\\subset \\partial M$ on which $d$ is finite. </p>\n\n<p>Then $M$ is obtained by taking $U=\\{ (x,t)\\subset \\partial M\\times [0,\\infty) | t\\leq d(x)\\}$ and taking the quotient of $graph(d_{|V})$ by the identification map to the cut locus.</p>\n\n<p>If the multiplicity $\\geq 2$ points are not dense in the cut locus, then there is an open subset in the cut locus of multiplicity one. This should pull back to an open subset of $graph(d_{|V})$, which maps homeomorophically under the identification map. But an open ball subset of $graph(d_{|V})$ mapping to the cut locus with multiplicity one will have a collar neighborhood homeomorphic to a half-space, which maps homeomorphically to $M$. However, the boundary of the half-space maps to the cut locus, not to $\\partial M$, which is a contradiction to the fact that $M$ is a manifold near the cut locus. </p>\n" }, { "answer_id": 261234, "score": 5, "text": "<p><em>Let me sketch another proof. It use analysis instead of topology. Namely we use existence of flow for continuous vector fields (uniqueness requires more, but we will not use it).</em></p>\n\n<p>Denote by <span class=\"math-container\">$K$</span> the subset in the interior of <span class=\"math-container\">$M$</span> with multiplicity 1\nand let <span class=\"math-container\">$U\\subset K$</span> be its interior. We need to show that <span class=\"math-container\">$U$</span> does not intersect the cut locus.</p>\n\n<p>Consider the function <span class=\"math-container\">$f= \\mathrm{dist}_{∂M}$</span>.\nNote that <span class=\"math-container\">$K$</span> is the set of points where <span class=\"math-container\">$f$</span> is differentiable, <span class=\"math-container\">$|\\nabla_x\\,f|=1$</span> for any <span class=\"math-container\">$x\\in K$</span> and <span class=\"math-container\">$\\nabla f$</span> is a continuous vector field in <span class=\"math-container\">$U$</span>. \nIn particular <span class=\"math-container\">$f$</span>-gradient flow exist in <span class=\"math-container\">$U$</span> (in fact, it is unique, but we do not need it).</p>\n\n<p>Fix <span class=\"math-container\">$x \\in U$</span>.\nMoving <span class=\"math-container\">$x$</span> along the <span class=\"math-container\">$f$</span>-gradient flow, we get a curve <span class=\"math-container\">$\\gamma$</span> such that <span class=\"math-container\">$(f\\circ\\gamma)'\\equiv 1$</span>.\nIn other words, <span class=\"math-container\">$\\gamma$</span> is an extension of the minimizing geodesic from <span class=\"math-container\">$∂M$</span> to <span class=\"math-container\">$x$</span>.\nHence <span class=\"math-container\">$x$</span> does not lie on the cut locus.</p>\n\n<p>Since <span class=\"math-container\">$x\\in U$</span> is arbitrary, the statement follows.</p>\n" }, { "answer_id": 261914, "score": 5, "text": "<p><em>I've got a letter from Stephanie Alexander with a complete answer. Let me summarize it here.</em></p>\n\n<p>The first proof is given in 4.8 of \"Schnittort und konvexe Mengen...\" by Hermann Karcher (1968). (The formulation is slightly weaker, but from the proof proves our statement follows; the idea is the same as in the answer of Ian Agol.)</p>\n\n<p>An other proof was given in \"Decomposition of cut loci\" by Richard Bishop (1977).</p>\n\n<p>Latter it appears as Lemma 2 in \"Distance function and cut loci...\" by Franz-Erich Wolter (1979). The proof is basically\nthe same as Karcher's.</p>\n" }, { "answer_id": 260949, "score": 6, "text": "<p>Intuitively, this seems plausible if $M$ is complete. Let $d:\\partial M \\to [0,\\infty]$ be the distance function to the cut locus. Then $d$ is continuous (by completeness), and $graph(d) \\subset \\partial M\\times [0,\\infty]$ is a copy of $\\partial M$. There is an open subset $V\\subset \\partial M$ on which $d$ is finite. </p>\n\n<p>Then $M$ is obtained by taking $U=\\{ (x,t)\\subset \\partial M\\times [0,\\infty) | t\\leq d(x)\\}$ and taking the quotient of $graph(d_{|V})$ by the identification map to the cut locus.</p>\n\n<p>If the multiplicity $\\geq 2$ points are not dense in the cut locus, then there is an open subset in the cut locus of multiplicity one. This should pull back to an open subset of $graph(d_{|V})$, which maps homeomorophically under the identification map. But an open ball subset of $graph(d_{|V})$ mapping to the cut locus with multiplicity one will have a collar neighborhood homeomorphic to a half-space, which maps homeomorphically to $M$. However, the boundary of the half-space maps to the cut locus, not to $\\partial M$, which is a contradiction to the fact that $M$ is a manifold near the cut locus. </p>\n" }, { "answer_id": 261234, "score": 5, "text": "<p><em>Let me sketch another proof. It use analysis instead of topology. Namely we use existence of flow for continuous vector fields (uniqueness requires more, but we will not use it).</em></p>\n\n<p>Denote by <span class=\"math-container\">$K$</span> the subset in the interior of <span class=\"math-container\">$M$</span> with multiplicity 1\nand let <span class=\"math-container\">$U\\subset K$</span> be its interior. We need to show that <span class=\"math-container\">$U$</span> does not intersect the cut locus.</p>\n\n<p>Consider the function <span class=\"math-container\">$f= \\mathrm{dist}_{∂M}$</span>.\nNote that <span class=\"math-container\">$K$</span> is the set of points where <span class=\"math-container\">$f$</span> is differentiable, <span class=\"math-container\">$|\\nabla_x\\,f|=1$</span> for any <span class=\"math-container\">$x\\in K$</span> and <span class=\"math-container\">$\\nabla f$</span> is a continuous vector field in <span class=\"math-container\">$U$</span>. \nIn particular <span class=\"math-container\">$f$</span>-gradient flow exist in <span class=\"math-container\">$U$</span> (in fact, it is unique, but we do not need it).</p>\n\n<p>Fix <span class=\"math-container\">$x \\in U$</span>.\nMoving <span class=\"math-container\">$x$</span> along the <span class=\"math-container\">$f$</span>-gradient flow, we get a curve <span class=\"math-container\">$\\gamma$</span> such that <span class=\"math-container\">$(f\\circ\\gamma)'\\equiv 1$</span>.\nIn other words, <span class=\"math-container\">$\\gamma$</span> is an extension of the minimizing geodesic from <span class=\"math-container\">$∂M$</span> to <span class=\"math-container\">$x$</span>.\nHence <span class=\"math-container\">$x$</span> does not lie on the cut locus.</p>\n\n<p>Since <span class=\"math-container\">$x\\in U$</span> is arbitrary, the statement follows.</p>\n" }, { "answer_id": 261914, "score": 5, "text": "<p><em>I've got a letter from Stephanie Alexander with a complete answer. Let me summarize it here.</em></p>\n\n<p>The first proof is given in 4.8 of \"Schnittort und konvexe Mengen...\" by Hermann Karcher (1968). (The formulation is slightly weaker, but from the proof proves our statement follows; the idea is the same as in the answer of Ian Agol.)</p>\n\n<p>An other proof was given in \"Decomposition of cut loci\" by Richard Bishop (1977).</p>\n\n<p>Latter it appears as Lemma 2 in \"Distance function and cut loci...\" by Franz-Erich Wolter (1979). The proof is basically\nthe same as Karcher's.</p>\n" } ]
Question: Conjugate points on cut locus Answer 1: <p>Intuitively, this seems plausible if $M$ is complete. Let $d:\partial M \to [0,\infty]$ be the distance function to the cut locus. Then $d$ is continuous (by completeness), and $graph(d) \subset \partial M\times [0,\infty]$ is a copy of $\partial M$. There is an open subset $V\subset \partial M$ on which $d$ is finite. </p> <p>Then $M$ is obtained by taking $U=\{ (x,t)\subset \partial M\times [0,\infty) | t\leq d(x)\}$ and taking the quotient of $graph(d_{|V})$ by the identification map to the cut locus.</p> <p>If the multiplicity $\geq 2$ points are not dense in the cut locus, then there is an open subset in the cut locus of multiplicity one. This should pull back to an open subset of $graph(d_{|V})$, which maps homeomorophically under the identification map. But an open ball subset of $graph(d_{|V})$ mapping to the cut locus with multiplicity one will have a collar neighborhood homeomorphic to a half-space, which maps homeomorphically to $M$. However, the boundary of the half-space maps to the cut locus, not to $\partial M$, which is a contradiction to the fact that $M$ is a manifold near the cut locus. </p> Answer 2: <p><em>Let me sketch another proof. It use analysis instead of topology. Namely we use existence of flow for continuous vector fields (uniqueness requires more, but we will not use it).</em></p> <p>Denote by <span class="math-container">$K$</span> the subset in the interior of <span class="math-container">$M$</span> with multiplicity 1 and let <span class="math-container">$U\subset K$</span> be its interior. We need to show that <span class="math-container">$U$</span> does not intersect the cut locus.</p> <p>Consider the function <span class="math-container">$f= \mathrm{dist}_{∂M}$</span>. Note that <span class="math-container">$K$</span> is the set of points where <span class="math-container">$f$</span> is differentiable, <span class="math-container">$|\nabla_x\,f|=1$</span> for any <span class="math-container">$x\in K$</span> and <span class="math-container">$\nabla f$</span> is a continuous vector field in <span class="math-container">$U$</span>. In particular <span class="math-container">$f$</span>-gradient flow exist in <span class="math-container">$U$</span> (in fact, it is unique, but we do not need it).</p> <p>Fix <span class="math-container">$x \in U$</span>. Moving <span class="math-container">$x$</span> along the <span class="math-container">$f$</span>-gradient flow, we get a curve <span class="math-container">$\gamma$</span> such that <span class="math-container">$(f\circ\gamma)'\equiv 1$</span>. In other words, <span class="math-container">$\gamma$</span> is an extension of the minimizing geodesic from <span class="math-container">$∂M$</span> to <span class="math-container">$x$</span>. Hence <span class="math-container">$x$</span> does not lie on the cut locus.</p> <p>Since <span class="math-container">$x\in U$</span> is arbitrary, the statement follows.</p> Answer 3: <p><em>I've got a letter from Stephanie Alexander with a complete answer. Let me summarize it here.</em></p> <p>The first proof is given in 4.8 of "Schnittort und konvexe Mengen..." by Hermann Karcher (1968). (The formulation is slightly weaker, but from the proof proves our statement follows; the idea is the same as in the answer of Ian Agol.)</p> <p>An other proof was given in "Decomposition of cut loci" by Richard Bishop (1977).</p> <p>Latter it appears as Lemma 2 in "Distance function and cut loci..." by Franz-Erich Wolter (1979). The proof is basically the same as Karcher's.</p> Answer 4: <p>Intuitively, this seems plausible if $M$ is complete. Let $d:\partial M \to [0,\infty]$ be the distance function to the cut locus. Then $d$ is continuous (by completeness), and $graph(d) \subset \partial M\times [0,\infty]$ is a copy of $\partial M$. There is an open subset $V\subset \partial M$ on which $d$ is finite. </p> <p>Then $M$ is obtained by taking $U=\{ (x,t)\subset \partial M\times [0,\infty) | t\leq d(x)\}$ and taking the quotient of $graph(d_{|V})$ by the identification map to the cut locus.</p> <p>If the multiplicity $\geq 2$ points are not dense in the cut locus, then there is an open subset in the cut locus of multiplicity one. This should pull back to an open subset of $graph(d_{|V})$, which maps homeomorophically under the identification map. But an open ball subset of $graph(d_{|V})$ mapping to the cut locus with multiplicity one will have a collar neighborhood homeomorphic to a half-space, which maps homeomorphically to $M$. However, the boundary of the half-space maps to the cut locus, not to $\partial M$, which is a contradiction to the fact that $M$ is a manifold near the cut locus. </p> Answer 5: <p><em>Let me sketch another proof. It use analysis instead of topology. Namely we use existence of flow for continuous vector fields (uniqueness requires more, but we will not use it).</em></p> <p>Denote by <span class="math-container">$K$</span> the subset in the interior of <span class="math-container">$M$</span> with multiplicity 1 and let <span class="math-container">$U\subset K$</span> be its interior. We need to show that <span class="math-container">$U$</span> does not intersect the cut locus.</p> <p>Consider the function <span class="math-container">$f= \mathrm{dist}_{∂M}$</span>. Note that <span class="math-container">$K$</span> is the set of points where <span class="math-container">$f$</span> is differentiable, <span class="math-container">$|\nabla_x\,f|=1$</span> for any <span class="math-container">$x\in K$</span> and <span class="math-container">$\nabla f$</span> is a continuous vector field in <span class="math-container">$U$</span>. In particular <span class="math-container">$f$</span>-gradient flow exist in <span class="math-container">$U$</span> (in fact, it is unique, but we do not need it).</p> <p>Fix <span class="math-container">$x \in U$</span>. Moving <span class="math-container">$x$</span> along the <span class="math-container">$f$</span>-gradient flow, we get a curve <span class="math-container">$\gamma$</span> such that <span class="math-container">$(f\circ\gamma)'\equiv 1$</span>. In other words, <span class="math-container">$\gamma$</span> is an extension of the minimizing geodesic from <span class="math-container">$∂M$</span> to <span class="math-container">$x$</span>. Hence <span class="math-container">$x$</span> does not lie on the cut locus.</p> <p>Since <span class="math-container">$x\in U$</span> is arbitrary, the statement follows.</p> Answer 6: <p><em>I've got a letter from Stephanie Alexander with a complete answer. Let me summarize it here.</em></p> <p>The first proof is given in 4.8 of "Schnittort und konvexe Mengen..." by Hermann Karcher (1968). (The formulation is slightly weaker, but from the proof proves our statement follows; the idea is the same as in the answer of Ian Agol.)</p> <p>An other proof was given in "Decomposition of cut loci" by Richard Bishop (1977).</p> <p>Latter it appears as Lemma 2 in "Distance function and cut loci..." by Franz-Erich Wolter (1979). The proof is basically the same as Karcher's.</p> Answer 7: <p>Intuitively, this seems plausible if $M$ is complete. Let $d:\partial M \to [0,\infty]$ be the distance function to the cut locus. Then $d$ is continuous (by completeness), and $graph(d) \subset \partial M\times [0,\infty]$ is a copy of $\partial M$. There is an open subset $V\subset \partial M$ on which $d$ is finite. </p> <p>Then $M$ is obtained by taking $U=\{ (x,t)\subset \partial M\times [0,\infty) | t\leq d(x)\}$ and taking the quotient of $graph(d_{|V})$ by the identification map to the cut locus.</p> <p>If the multiplicity $\geq 2$ points are not dense in the cut locus, then there is an open subset in the cut locus of multiplicity one. This should pull back to an open subset of $graph(d_{|V})$, which maps homeomorophically under the identification map. But an open ball subset of $graph(d_{|V})$ mapping to the cut locus with multiplicity one will have a collar neighborhood homeomorphic to a half-space, which maps homeomorphically to $M$. However, the boundary of the half-space maps to the cut locus, not to $\partial M$, which is a contradiction to the fact that $M$ is a manifold near the cut locus. </p> Answer 8: <p><em>Let me sketch another proof. It use analysis instead of topology. Namely we use existence of flow for continuous vector fields (uniqueness requires more, but we will not use it).</em></p> <p>Denote by <span class="math-container">$K$</span> the subset in the interior of <span class="math-container">$M$</span> with multiplicity 1 and let <span class="math-container">$U\subset K$</span> be its interior. We need to show that <span class="math-container">$U$</span> does not intersect the cut locus.</p> <p>Consider the function <span class="math-container">$f= \mathrm{dist}_{∂M}$</span>. Note that <span class="math-container">$K$</span> is the set of points where <span class="math-container">$f$</span> is differentiable, <span class="math-container">$|\nabla_x\,f|=1$</span> for any <span class="math-container">$x\in K$</span> and <span class="math-container">$\nabla f$</span> is a continuous vector field in <span class="math-container">$U$</span>. In particular <span class="math-container">$f$</span>-gradient flow exist in <span class="math-container">$U$</span> (in fact, it is unique, but we do not need it).</p> <p>Fix <span class="math-container">$x \in U$</span>. Moving <span class="math-container">$x$</span> along the <span class="math-container">$f$</span>-gradient flow, we get a curve <span class="math-container">$\gamma$</span> such that <span class="math-container">$(f\circ\gamma)'\equiv 1$</span>. In other words, <span class="math-container">$\gamma$</span> is an extension of the minimizing geodesic from <span class="math-container">$∂M$</span> to <span class="math-container">$x$</span>. Hence <span class="math-container">$x$</span> does not lie on the cut locus.</p> <p>Since <span class="math-container">$x\in U$</span> is arbitrary, the statement follows.</p> Answer 9: <p><em>I've got a letter from Stephanie Alexander with a complete answer. Let me summarize it here.</em></p> <p>The first proof is given in 4.8 of "Schnittort und konvexe Mengen..." by Hermann Karcher (1968). (The formulation is slightly weaker, but from the proof proves our statement follows; the idea is the same as in the answer of Ian Agol.)</p> <p>An other proof was given in "Decomposition of cut loci" by Richard Bishop (1977).</p> <p>Latter it appears as Lemma 2 in "Distance function and cut loci..." by Franz-Erich Wolter (1979). The proof is basically the same as Karcher's.</p>
Can there be a polymath project for mathematical physics?
260,511
14
https://mathoverflow.net/questions/260511/can-there-be-a-polymath-project-for-mathematical-physics
<p>My hunch is that it might be possible to create something like <a href="https://polymathprojects.org/" rel="noreferrer">https://polymathprojects.org/</a> for mathematical physics and I'd like to know whether MathOverflow users can recommend some appropriate problems. As Abdelmalek Abdessalam pointed out in the comments below such a platform can potentially be used to develop research programs in mathematical physics that are suitable for polymath type collaborative work.</p> <p>The gist of this project would be to use collaborate problem-solving to tackle open problems in mathematical physics. Timothy Gowers wrote a seminal blog post on the subject here: <a href="https://gowers.wordpress.com/2009/01/27/is-massively-collaborative-mathematics-possible/" rel="noreferrer">Is massively collaborative mathematics possible?</a></p> <p>There's a good point made by a commenter below as to why not simply use the current polymath platform. Well, I think that it would be easier for such a platform to succeed and gain support among mathematical physicists if it was maintained by mathematical physicists and focused mainly on problems in the area of mathematical physics. This is not the current setup of the <a href="https://polymathprojects.org/" rel="noreferrer">Polymath project</a>. </p>
[ { "answer_id": 260583, "score": 4, "text": "<p>Polymath 7 tackled the hot spots conjecture(due to Jeffrey Rauch) in 2012: </p>\n\n<blockquote>\n <p>Suppose a flat piece of metal, represented by a two-dimensional\n bounded connected domain, is given an initial heat distribution which\n then flows throughout the metal. Assuming the metal is insulated (i.e.\n no heat escapes from the piece of metal), then given enough time, the\n hottest point on the metal will lie on its boundary.</p>\n</blockquote>\n\n<p>The goal of the project was to establish the conjecture for a large range of triangles and <a href=\"https://polymathprojects.org/2012/06/24/polymath7-research-threads-3-the-hot-spots-conjecture/\" rel=\"nofollow noreferrer\">partial results were documented on the blog</a>. I must note that the associated <a href=\"http://michaelnielsen.org/polymath1/index.php?title=The_hot_spots_conjecture\" rel=\"nofollow noreferrer\">wiki page</a> is also very useful as it lists different cases as well as different approaches. </p>\n\n<p>I think this is a nice example of a non-trivial problem that can potentially be tackled on a collaborative problem solving platform focused on mathematical physics. </p>\n" }, { "answer_id": 260745, "score": 2, "text": "<p>I recommend the following problem :</p>\n\n<p>Proof the existence of the spectral gap for the fractional quantum Hall effect at least for a simplified model where Laughlin's wavefunction is the exact ground state.</p>\n" }, { "answer_id": 341491, "score": 1, "text": "<p>Many problems in the IQOQI list of <a href=\"https://oqp.iqoqi.univie.ac.at/open-quantum-problems\" rel=\"nofollow noreferrer\">open quantum problems</a> have a strong aspect of mathematical physics, these could be of interest for such possible project.</p>\n\n<p>Addendum: The question of OP was \"I'd like to know whether MathOverflow users can recommend some appropriate problems.\" My answer does that, similar to the answer of user jjcale above.</p>\n" }, { "answer_id": 260583, "score": 4, "text": "<p>Polymath 7 tackled the hot spots conjecture(due to Jeffrey Rauch) in 2012: </p>\n\n<blockquote>\n <p>Suppose a flat piece of metal, represented by a two-dimensional\n bounded connected domain, is given an initial heat distribution which\n then flows throughout the metal. Assuming the metal is insulated (i.e.\n no heat escapes from the piece of metal), then given enough time, the\n hottest point on the metal will lie on its boundary.</p>\n</blockquote>\n\n<p>The goal of the project was to establish the conjecture for a large range of triangles and <a href=\"https://polymathprojects.org/2012/06/24/polymath7-research-threads-3-the-hot-spots-conjecture/\" rel=\"nofollow noreferrer\">partial results were documented on the blog</a>. I must note that the associated <a href=\"http://michaelnielsen.org/polymath1/index.php?title=The_hot_spots_conjecture\" rel=\"nofollow noreferrer\">wiki page</a> is also very useful as it lists different cases as well as different approaches. </p>\n\n<p>I think this is a nice example of a non-trivial problem that can potentially be tackled on a collaborative problem solving platform focused on mathematical physics. </p>\n" }, { "answer_id": 260745, "score": 2, "text": "<p>I recommend the following problem :</p>\n\n<p>Proof the existence of the spectral gap for the fractional quantum Hall effect at least for a simplified model where Laughlin's wavefunction is the exact ground state.</p>\n" }, { "answer_id": 341491, "score": 1, "text": "<p>Many problems in the IQOQI list of <a href=\"https://oqp.iqoqi.univie.ac.at/open-quantum-problems\" rel=\"nofollow noreferrer\">open quantum problems</a> have a strong aspect of mathematical physics, these could be of interest for such possible project.</p>\n\n<p>Addendum: The question of OP was \"I'd like to know whether MathOverflow users can recommend some appropriate problems.\" My answer does that, similar to the answer of user jjcale above.</p>\n" }, { "answer_id": 260583, "score": 4, "text": "<p>Polymath 7 tackled the hot spots conjecture(due to Jeffrey Rauch) in 2012: </p>\n\n<blockquote>\n <p>Suppose a flat piece of metal, represented by a two-dimensional\n bounded connected domain, is given an initial heat distribution which\n then flows throughout the metal. Assuming the metal is insulated (i.e.\n no heat escapes from the piece of metal), then given enough time, the\n hottest point on the metal will lie on its boundary.</p>\n</blockquote>\n\n<p>The goal of the project was to establish the conjecture for a large range of triangles and <a href=\"https://polymathprojects.org/2012/06/24/polymath7-research-threads-3-the-hot-spots-conjecture/\" rel=\"nofollow noreferrer\">partial results were documented on the blog</a>. I must note that the associated <a href=\"http://michaelnielsen.org/polymath1/index.php?title=The_hot_spots_conjecture\" rel=\"nofollow noreferrer\">wiki page</a> is also very useful as it lists different cases as well as different approaches. </p>\n\n<p>I think this is a nice example of a non-trivial problem that can potentially be tackled on a collaborative problem solving platform focused on mathematical physics. </p>\n" }, { "answer_id": 260745, "score": 2, "text": "<p>I recommend the following problem :</p>\n\n<p>Proof the existence of the spectral gap for the fractional quantum Hall effect at least for a simplified model where Laughlin's wavefunction is the exact ground state.</p>\n" }, { "answer_id": 341491, "score": 1, "text": "<p>Many problems in the IQOQI list of <a href=\"https://oqp.iqoqi.univie.ac.at/open-quantum-problems\" rel=\"nofollow noreferrer\">open quantum problems</a> have a strong aspect of mathematical physics, these could be of interest for such possible project.</p>\n\n<p>Addendum: The question of OP was \"I'd like to know whether MathOverflow users can recommend some appropriate problems.\" My answer does that, similar to the answer of user jjcale above.</p>\n" } ]
Question: Can there be a polymath project for mathematical physics? Answer 1: <p>Polymath 7 tackled the hot spots conjecture(due to Jeffrey Rauch) in 2012: </p> <blockquote> <p>Suppose a flat piece of metal, represented by a two-dimensional bounded connected domain, is given an initial heat distribution which then flows throughout the metal. Assuming the metal is insulated (i.e. no heat escapes from the piece of metal), then given enough time, the hottest point on the metal will lie on its boundary.</p> </blockquote> <p>The goal of the project was to establish the conjecture for a large range of triangles and <a href="https://polymathprojects.org/2012/06/24/polymath7-research-threads-3-the-hot-spots-conjecture/" rel="nofollow noreferrer">partial results were documented on the blog</a>. I must note that the associated <a href="http://michaelnielsen.org/polymath1/index.php?title=The_hot_spots_conjecture" rel="nofollow noreferrer">wiki page</a> is also very useful as it lists different cases as well as different approaches. </p> <p>I think this is a nice example of a non-trivial problem that can potentially be tackled on a collaborative problem solving platform focused on mathematical physics. </p> Answer 2: <p>I recommend the following problem :</p> <p>Proof the existence of the spectral gap for the fractional quantum Hall effect at least for a simplified model where Laughlin's wavefunction is the exact ground state.</p> Answer 3: <p>Many problems in the IQOQI list of <a href="https://oqp.iqoqi.univie.ac.at/open-quantum-problems" rel="nofollow noreferrer">open quantum problems</a> have a strong aspect of mathematical physics, these could be of interest for such possible project.</p> <p>Addendum: The question of OP was "I'd like to know whether MathOverflow users can recommend some appropriate problems." My answer does that, similar to the answer of user jjcale above.</p> Answer 4: <p>Polymath 7 tackled the hot spots conjecture(due to Jeffrey Rauch) in 2012: </p> <blockquote> <p>Suppose a flat piece of metal, represented by a two-dimensional bounded connected domain, is given an initial heat distribution which then flows throughout the metal. Assuming the metal is insulated (i.e. no heat escapes from the piece of metal), then given enough time, the hottest point on the metal will lie on its boundary.</p> </blockquote> <p>The goal of the project was to establish the conjecture for a large range of triangles and <a href="https://polymathprojects.org/2012/06/24/polymath7-research-threads-3-the-hot-spots-conjecture/" rel="nofollow noreferrer">partial results were documented on the blog</a>. I must note that the associated <a href="http://michaelnielsen.org/polymath1/index.php?title=The_hot_spots_conjecture" rel="nofollow noreferrer">wiki page</a> is also very useful as it lists different cases as well as different approaches. </p> <p>I think this is a nice example of a non-trivial problem that can potentially be tackled on a collaborative problem solving platform focused on mathematical physics. </p> Answer 5: <p>I recommend the following problem :</p> <p>Proof the existence of the spectral gap for the fractional quantum Hall effect at least for a simplified model where Laughlin's wavefunction is the exact ground state.</p> Answer 6: <p>Many problems in the IQOQI list of <a href="https://oqp.iqoqi.univie.ac.at/open-quantum-problems" rel="nofollow noreferrer">open quantum problems</a> have a strong aspect of mathematical physics, these could be of interest for such possible project.</p> <p>Addendum: The question of OP was "I'd like to know whether MathOverflow users can recommend some appropriate problems." My answer does that, similar to the answer of user jjcale above.</p> Answer 7: <p>Polymath 7 tackled the hot spots conjecture(due to Jeffrey Rauch) in 2012: </p> <blockquote> <p>Suppose a flat piece of metal, represented by a two-dimensional bounded connected domain, is given an initial heat distribution which then flows throughout the metal. Assuming the metal is insulated (i.e. no heat escapes from the piece of metal), then given enough time, the hottest point on the metal will lie on its boundary.</p> </blockquote> <p>The goal of the project was to establish the conjecture for a large range of triangles and <a href="https://polymathprojects.org/2012/06/24/polymath7-research-threads-3-the-hot-spots-conjecture/" rel="nofollow noreferrer">partial results were documented on the blog</a>. I must note that the associated <a href="http://michaelnielsen.org/polymath1/index.php?title=The_hot_spots_conjecture" rel="nofollow noreferrer">wiki page</a> is also very useful as it lists different cases as well as different approaches. </p> <p>I think this is a nice example of a non-trivial problem that can potentially be tackled on a collaborative problem solving platform focused on mathematical physics. </p> Answer 8: <p>I recommend the following problem :</p> <p>Proof the existence of the spectral gap for the fractional quantum Hall effect at least for a simplified model where Laughlin's wavefunction is the exact ground state.</p> Answer 9: <p>Many problems in the IQOQI list of <a href="https://oqp.iqoqi.univie.ac.at/open-quantum-problems" rel="nofollow noreferrer">open quantum problems</a> have a strong aspect of mathematical physics, these could be of interest for such possible project.</p> <p>Addendum: The question of OP was "I'd like to know whether MathOverflow users can recommend some appropriate problems." My answer does that, similar to the answer of user jjcale above.</p>
Expected number of lines meeting four given lines or &quot;what is 1.72...&quot;
260,607
14
https://mathoverflow.net/questions/260607/expected-number-of-lines-meeting-four-given-lines-or-what-is-1-72
<p>Given four random lines in <span class="math-container">$\mathbb{R}P^3$</span>, how many lines intersect all of those lines? In the recent paper <a href="http://arxiv.org/abs/1612.06893v1" rel="nofollow noreferrer">Probabilistic Schubert Calculus</a>, Peter Bürgisser and Antonio Lerario discuss this question and much more general versions of it.</p> <p>In <strong>Proposition 6.7</strong>, they determine the expected number of lines meeting four random lines in <span class="math-container">$\mathbb{R}P^3$</span> to be:</p> <p><span class="math-container">$$\operatorname{edeg}G(2,4) =\\ 2^{-13}\int_{[0,2\pi]^6}\left| \det\begin{pmatrix} \sin{t_1}\sin{s_1}&amp;\sin{t_2}\sin{s_2}&amp;\sin{t_3}\sin{s_3}\\ \cos{t_1}\sin{s_1}&amp;\cos{t_2}\sin{s_2}&amp;\cos{t_3}\sin{s_3}\\ \sin{t_1}\cos{s_1}&amp;\sin{t_2}\cos{s_2}&amp;\sin{t_3}\cos{s_3} \end{pmatrix}\right|dt_1dt_2dt_3ds_1ds_2ds_3$$</span></p> <p>The integrand can be expanded to <span class="math-container">$$\left|\cos(s_2)\sin(s_1)\sin(s_3)\sin(t_2)\sin(t_1 - t_3)- \sin(s_2)\big(\cos(s_1)\sin(s_3)\sin(t_1)\sin(t_2 - t_3) + \cos(s_3) \sin(s_1)\sin(t_3)\sin(t_1 - t_2)\big)\right|$$</span></p> <p>The absolute value in the integrand and the integration over six variables seem to make it difficult to evaluate this integral to high precision. I am interested in the exact value of this number. Bürgisser and Lerario computed it to be <span class="math-container">$1.72$</span> rounded to two digits; I ran <span class="math-container">$10^{11}$</span> Monte Carlo evaluations to obtain <span class="math-container">$1.726225\dots$</span> with an estimated error of <span class="math-container">$7.3\cdot 10^{-6}$</span></p> <p>My questions are:</p> <ul> <li>Can the exact value of <span class="math-container">$\operatorname{edeg}G(2,4)$</span> be determined; is there perhaps a relation with other constants; is it, for example, algebraic over <span class="math-container">$\mathbb{Q}[\pi]$</span>?</li> <li>If this is too difficult, what are ways of evaluating the above integral to higher accuracy numerically?</li> </ul> <p><strong>Update</strong>: In AdamP.Goucher's great answer, he provides more digits (<span class="math-container">$1.726230876$</span>) and also uses a reformulation described by Matt F. in a comment to get an integrand without any absolute values:</p> <p><span class="math-container">$$G(2,4) = 2^{-6} \int \dfrac{h\left((y-u)^2+(v-x)^2\right)^2}{\left((1+(u + (y-u)c - (v-x)h)^2)(1+(x + (v-x)c + (y-u)h)^2)\prod_{\alpha\in\{u,y,x,v\}}(1+\alpha^2)\right)^{3/2}}$$</span> where we integrate over <span class="math-container">$\mathbb{R}^5\times \mathbb{R}^+$</span> (and <span class="math-container">$h$</span> is the positive variable).</p> <p>Perhaps this form could come in handy for evaluating this integral to higher accuracy?!</p> <hr /> <p><strong>Update 2</strong> The new formula from L.Mathis and Antonio Lerario is very useful for calculating digits! The following <a href="http://mpmath.org/" rel="nofollow noreferrer">mpmath</a> code can returns in less than <span class="math-container">$3$</span> minutes</p> <p><span class="math-container">$1.7262312489219034885256331685361697650475579915479447$</span></p> <p>(the last few digits might not be accurate) I expect to make that even much faster when solving the two integrals <span class="math-container">$F$</span> and <span class="math-container">$G$</span> symbolically first.</p> <pre><code>import functools from mpmath import mp @functools.lru_cache(maxsize=1000) def F(u): return mp.quad(lambda phi: (u*mp.sin(phi)**2)/(mp.sqrt(mp.cos(phi)**2 + u**2*mp.sin(phi)**2)), [0, mp.pi/2]) @functools.lru_cache(maxsize=1000) def G(u): return mp.quad(lambda phi: (mp.sin(phi)**2)/(mp.sqrt(mp.sin(phi)**2 + u**2*mp.cos(phi)**2)), [0, mp.pi/2]) def H(u): return F(u)/G(u) def L(u): return F(u)*G(u) def integrand3(u): return L(u)**2*(1/H(u) - H(u))*mp.diff(H, u)/H(u) dps = 50 mp.dps = dps %time z = 3*mp.quad(integrand, [0,1]); z </code></pre>
[ { "answer_id": 260719, "score": 11, "text": "<p>The integrand is periodic modulo $\\pi$ in each variable, so it suffices to integrate each variable over $[0, \\pi]$ and replace the constant factor by $2^{-7}$.</p>\n\n<p>If we were to apply a change of variables (e.g. set $x = \\cos(s_1)$ and similarly for the other five variables), we would have an integral of a piecewise-algebraic function which thus belongs to Kontsevich and Zagier's 'ring of periods'. Alas, there is no proven algorithm for determining whether periods are expressible in terms of elementary functions.</p>\n\n<p>Now, using Gauss-Legendre integration with $N$ points in each variable, we can approximate it with a weighted sum of $N^6$ evaluations of the integrand. Since the absolute value of the determinant is unchanged under permuting the columns, and zero whenever two columns are equal, we can reduce this to $\\binom{N^2}{3}$ function evaluations.</p>\n\n<p>I managed to implement this in C code, where each iteration, amortized, only takes four floating-point additions, three multiplications, and an absolute-value calculation. With several hours running on a 144-core machine, it could do the calculation for both $N = 401$ and $N = 409$. The results were 1.726230867 and 1.726230885, respectively.</p>\n\n<p>Consequently, the integral is roughly 1.726230876, with an expected error on the order of $10^{-8}$.</p>\n\n<hr>\n\n<p><strong>EDIT:</strong> Matt F. posted an algebraic form for the integral, by taking $u = \\cot(t1)$, $v = \\cot(s2)$, $w = \\cot(t3)$, $x = \\cot(s1)$, $y = \\cot(t2)$, $z = \\cot(s3)$:</p>\n\n<p>$$ I = 2^{-7} \\int \\dfrac{|uv-vw+wx-xy+yz-zu|}{[(1+u^2)(1+v^2)(1+w^2)(1+x^2)(1+y^2)(1+z^2)]^{3/2}} $$</p>\n\n<p>where the integral is taken over $(u,v,w,x,y,z) \\in \\mathbb{R}^6$. Observing that the numerator is twice the area of a triangle with vertices $(u, x), (y, v), (w, z)$, we can perform a change of variables by setting:</p>\n\n<p>$$ w = u + (y-u)c - (v-x)h \\textrm{ ; } z = x + (v-x)c + (y-u)h $$</p>\n\n<p>where $c \\in \\mathbb{R}$ and $h \\in \\mathbb{R}^{+}$. Taking the Jacobian into account, the integral becomes:</p>\n\n<p>$$ I = 2^{-6} \\int \\dfrac{h[(y-u)^2+(v-x)^2]^2}{[(1+u^2)(1+y^2)(1+(u + (y-u)c - (v-x)h)^2)(1+x^2)(1+v^2)(1+(x + (v-x)c + (y-u)h)^2)]^{3/2}} $$</p>\n\n<p>where $(u,y,x,v,c,h) \\in \\mathbb{R}^5 \\times \\mathbb{R}^+$. Although the integral is uglier, it satisfyingly has no absolute value operator.</p>\n" }, { "answer_id": 375630, "score": 4, "text": "<p>If anyone is still interested, Antonio Lerario and I recently published a paper: <a href=\"https://link.springer.com/article/10.1007/s40598-020-00160-w\" rel=\"nofollow noreferrer\">Probabilistic Schubert Calculus: Asymptotics</a>,\nin which we give a more convenient formula to compute this number.</p>\n<p>What was called before <span class=\"math-container\">$\\operatorname{edeg}G(2,4)$</span> is now denoted by <span class=\"math-container\">$\\delta_{1,3}$</span> and we give a line integral formula in Proposition 24:</p>\n<p><span class=\"math-container\">\\begin{aligned} \n\\delta _{1,3}=-6 \\pi ^{4}\\int _{0}^1 L(u)^{2}\\mathrm {sinh}(w(u))w'(u)\\mathrm{d}u \\end{aligned}</span>\nwhere <span class=\"math-container\">$L=F⋅G$</span> and <span class=\"math-container\">$w=log(F/G)$</span> with\n<span class=\"math-container\">\\begin{aligned} \nF(u)&amp;:=\\int _{0}^{\\pi /2} \\frac{u \\ \\sin ^2(\\theta )}{\\sqrt{\\cos ^2\\theta +u^2 \\sin ^2\\theta }}\\mathrm {d}\\theta \\\\ G (u)&amp;:=\\int _{0}^{\\pi /2} \\frac{\\sin ^2(\\theta )}{\\sqrt{\\sin ^2\\theta +u^2 \\cos ^2\\theta }}\\mathrm {d}\\theta .\n\\end{aligned}</span></p>\n<p>We didn't run any advanced numerical evaluation but this looks much nicer than the previous formula. I hope this will help. It is still unknown to us if this number can be expressed as a closed formula using special functions.</p>\n" }, { "answer_id": 260719, "score": 11, "text": "<p>The integrand is periodic modulo $\\pi$ in each variable, so it suffices to integrate each variable over $[0, \\pi]$ and replace the constant factor by $2^{-7}$.</p>\n\n<p>If we were to apply a change of variables (e.g. set $x = \\cos(s_1)$ and similarly for the other five variables), we would have an integral of a piecewise-algebraic function which thus belongs to Kontsevich and Zagier's 'ring of periods'. Alas, there is no proven algorithm for determining whether periods are expressible in terms of elementary functions.</p>\n\n<p>Now, using Gauss-Legendre integration with $N$ points in each variable, we can approximate it with a weighted sum of $N^6$ evaluations of the integrand. Since the absolute value of the determinant is unchanged under permuting the columns, and zero whenever two columns are equal, we can reduce this to $\\binom{N^2}{3}$ function evaluations.</p>\n\n<p>I managed to implement this in C code, where each iteration, amortized, only takes four floating-point additions, three multiplications, and an absolute-value calculation. With several hours running on a 144-core machine, it could do the calculation for both $N = 401$ and $N = 409$. The results were 1.726230867 and 1.726230885, respectively.</p>\n\n<p>Consequently, the integral is roughly 1.726230876, with an expected error on the order of $10^{-8}$.</p>\n\n<hr>\n\n<p><strong>EDIT:</strong> Matt F. posted an algebraic form for the integral, by taking $u = \\cot(t1)$, $v = \\cot(s2)$, $w = \\cot(t3)$, $x = \\cot(s1)$, $y = \\cot(t2)$, $z = \\cot(s3)$:</p>\n\n<p>$$ I = 2^{-7} \\int \\dfrac{|uv-vw+wx-xy+yz-zu|}{[(1+u^2)(1+v^2)(1+w^2)(1+x^2)(1+y^2)(1+z^2)]^{3/2}} $$</p>\n\n<p>where the integral is taken over $(u,v,w,x,y,z) \\in \\mathbb{R}^6$. Observing that the numerator is twice the area of a triangle with vertices $(u, x), (y, v), (w, z)$, we can perform a change of variables by setting:</p>\n\n<p>$$ w = u + (y-u)c - (v-x)h \\textrm{ ; } z = x + (v-x)c + (y-u)h $$</p>\n\n<p>where $c \\in \\mathbb{R}$ and $h \\in \\mathbb{R}^{+}$. Taking the Jacobian into account, the integral becomes:</p>\n\n<p>$$ I = 2^{-6} \\int \\dfrac{h[(y-u)^2+(v-x)^2]^2}{[(1+u^2)(1+y^2)(1+(u + (y-u)c - (v-x)h)^2)(1+x^2)(1+v^2)(1+(x + (v-x)c + (y-u)h)^2)]^{3/2}} $$</p>\n\n<p>where $(u,y,x,v,c,h) \\in \\mathbb{R}^5 \\times \\mathbb{R}^+$. Although the integral is uglier, it satisfyingly has no absolute value operator.</p>\n" }, { "answer_id": 375630, "score": 4, "text": "<p>If anyone is still interested, Antonio Lerario and I recently published a paper: <a href=\"https://link.springer.com/article/10.1007/s40598-020-00160-w\" rel=\"nofollow noreferrer\">Probabilistic Schubert Calculus: Asymptotics</a>,\nin which we give a more convenient formula to compute this number.</p>\n<p>What was called before <span class=\"math-container\">$\\operatorname{edeg}G(2,4)$</span> is now denoted by <span class=\"math-container\">$\\delta_{1,3}$</span> and we give a line integral formula in Proposition 24:</p>\n<p><span class=\"math-container\">\\begin{aligned} \n\\delta _{1,3}=-6 \\pi ^{4}\\int _{0}^1 L(u)^{2}\\mathrm {sinh}(w(u))w'(u)\\mathrm{d}u \\end{aligned}</span>\nwhere <span class=\"math-container\">$L=F⋅G$</span> and <span class=\"math-container\">$w=log(F/G)$</span> with\n<span class=\"math-container\">\\begin{aligned} \nF(u)&amp;:=\\int _{0}^{\\pi /2} \\frac{u \\ \\sin ^2(\\theta )}{\\sqrt{\\cos ^2\\theta +u^2 \\sin ^2\\theta }}\\mathrm {d}\\theta \\\\ G (u)&amp;:=\\int _{0}^{\\pi /2} \\frac{\\sin ^2(\\theta )}{\\sqrt{\\sin ^2\\theta +u^2 \\cos ^2\\theta }}\\mathrm {d}\\theta .\n\\end{aligned}</span></p>\n<p>We didn't run any advanced numerical evaluation but this looks much nicer than the previous formula. I hope this will help. It is still unknown to us if this number can be expressed as a closed formula using special functions.</p>\n" }, { "answer_id": 260719, "score": 11, "text": "<p>The integrand is periodic modulo $\\pi$ in each variable, so it suffices to integrate each variable over $[0, \\pi]$ and replace the constant factor by $2^{-7}$.</p>\n\n<p>If we were to apply a change of variables (e.g. set $x = \\cos(s_1)$ and similarly for the other five variables), we would have an integral of a piecewise-algebraic function which thus belongs to Kontsevich and Zagier's 'ring of periods'. Alas, there is no proven algorithm for determining whether periods are expressible in terms of elementary functions.</p>\n\n<p>Now, using Gauss-Legendre integration with $N$ points in each variable, we can approximate it with a weighted sum of $N^6$ evaluations of the integrand. Since the absolute value of the determinant is unchanged under permuting the columns, and zero whenever two columns are equal, we can reduce this to $\\binom{N^2}{3}$ function evaluations.</p>\n\n<p>I managed to implement this in C code, where each iteration, amortized, only takes four floating-point additions, three multiplications, and an absolute-value calculation. With several hours running on a 144-core machine, it could do the calculation for both $N = 401$ and $N = 409$. The results were 1.726230867 and 1.726230885, respectively.</p>\n\n<p>Consequently, the integral is roughly 1.726230876, with an expected error on the order of $10^{-8}$.</p>\n\n<hr>\n\n<p><strong>EDIT:</strong> Matt F. posted an algebraic form for the integral, by taking $u = \\cot(t1)$, $v = \\cot(s2)$, $w = \\cot(t3)$, $x = \\cot(s1)$, $y = \\cot(t2)$, $z = \\cot(s3)$:</p>\n\n<p>$$ I = 2^{-7} \\int \\dfrac{|uv-vw+wx-xy+yz-zu|}{[(1+u^2)(1+v^2)(1+w^2)(1+x^2)(1+y^2)(1+z^2)]^{3/2}} $$</p>\n\n<p>where the integral is taken over $(u,v,w,x,y,z) \\in \\mathbb{R}^6$. Observing that the numerator is twice the area of a triangle with vertices $(u, x), (y, v), (w, z)$, we can perform a change of variables by setting:</p>\n\n<p>$$ w = u + (y-u)c - (v-x)h \\textrm{ ; } z = x + (v-x)c + (y-u)h $$</p>\n\n<p>where $c \\in \\mathbb{R}$ and $h \\in \\mathbb{R}^{+}$. Taking the Jacobian into account, the integral becomes:</p>\n\n<p>$$ I = 2^{-6} \\int \\dfrac{h[(y-u)^2+(v-x)^2]^2}{[(1+u^2)(1+y^2)(1+(u + (y-u)c - (v-x)h)^2)(1+x^2)(1+v^2)(1+(x + (v-x)c + (y-u)h)^2)]^{3/2}} $$</p>\n\n<p>where $(u,y,x,v,c,h) \\in \\mathbb{R}^5 \\times \\mathbb{R}^+$. Although the integral is uglier, it satisfyingly has no absolute value operator.</p>\n" }, { "answer_id": 375630, "score": 4, "text": "<p>If anyone is still interested, Antonio Lerario and I recently published a paper: <a href=\"https://link.springer.com/article/10.1007/s40598-020-00160-w\" rel=\"nofollow noreferrer\">Probabilistic Schubert Calculus: Asymptotics</a>,\nin which we give a more convenient formula to compute this number.</p>\n<p>What was called before <span class=\"math-container\">$\\operatorname{edeg}G(2,4)$</span> is now denoted by <span class=\"math-container\">$\\delta_{1,3}$</span> and we give a line integral formula in Proposition 24:</p>\n<p><span class=\"math-container\">\\begin{aligned} \n\\delta _{1,3}=-6 \\pi ^{4}\\int _{0}^1 L(u)^{2}\\mathrm {sinh}(w(u))w'(u)\\mathrm{d}u \\end{aligned}</span>\nwhere <span class=\"math-container\">$L=F⋅G$</span> and <span class=\"math-container\">$w=log(F/G)$</span> with\n<span class=\"math-container\">\\begin{aligned} \nF(u)&amp;:=\\int _{0}^{\\pi /2} \\frac{u \\ \\sin ^2(\\theta )}{\\sqrt{\\cos ^2\\theta +u^2 \\sin ^2\\theta }}\\mathrm {d}\\theta \\\\ G (u)&amp;:=\\int _{0}^{\\pi /2} \\frac{\\sin ^2(\\theta )}{\\sqrt{\\sin ^2\\theta +u^2 \\cos ^2\\theta }}\\mathrm {d}\\theta .\n\\end{aligned}</span></p>\n<p>We didn't run any advanced numerical evaluation but this looks much nicer than the previous formula. I hope this will help. It is still unknown to us if this number can be expressed as a closed formula using special functions.</p>\n" } ]
Question: Expected number of lines meeting four given lines or &quot;what is 1.72...&quot; Answer 1: <p>The integrand is periodic modulo $\pi$ in each variable, so it suffices to integrate each variable over $[0, \pi]$ and replace the constant factor by $2^{-7}$.</p> <p>If we were to apply a change of variables (e.g. set $x = \cos(s_1)$ and similarly for the other five variables), we would have an integral of a piecewise-algebraic function which thus belongs to Kontsevich and Zagier's 'ring of periods'. Alas, there is no proven algorithm for determining whether periods are expressible in terms of elementary functions.</p> <p>Now, using Gauss-Legendre integration with $N$ points in each variable, we can approximate it with a weighted sum of $N^6$ evaluations of the integrand. Since the absolute value of the determinant is unchanged under permuting the columns, and zero whenever two columns are equal, we can reduce this to $\binom{N^2}{3}$ function evaluations.</p> <p>I managed to implement this in C code, where each iteration, amortized, only takes four floating-point additions, three multiplications, and an absolute-value calculation. With several hours running on a 144-core machine, it could do the calculation for both $N = 401$ and $N = 409$. The results were 1.726230867 and 1.726230885, respectively.</p> <p>Consequently, the integral is roughly 1.726230876, with an expected error on the order of $10^{-8}$.</p> <hr> <p><strong>EDIT:</strong> Matt F. posted an algebraic form for the integral, by taking $u = \cot(t1)$, $v = \cot(s2)$, $w = \cot(t3)$, $x = \cot(s1)$, $y = \cot(t2)$, $z = \cot(s3)$:</p> <p>$$ I = 2^{-7} \int \dfrac{|uv-vw+wx-xy+yz-zu|}{[(1+u^2)(1+v^2)(1+w^2)(1+x^2)(1+y^2)(1+z^2)]^{3/2}} $$</p> <p>where the integral is taken over $(u,v,w,x,y,z) \in \mathbb{R}^6$. Observing that the numerator is twice the area of a triangle with vertices $(u, x), (y, v), (w, z)$, we can perform a change of variables by setting:</p> <p>$$ w = u + (y-u)c - (v-x)h \textrm{ ; } z = x + (v-x)c + (y-u)h $$</p> <p>where $c \in \mathbb{R}$ and $h \in \mathbb{R}^{+}$. Taking the Jacobian into account, the integral becomes:</p> <p>$$ I = 2^{-6} \int \dfrac{h[(y-u)^2+(v-x)^2]^2}{[(1+u^2)(1+y^2)(1+(u + (y-u)c - (v-x)h)^2)(1+x^2)(1+v^2)(1+(x + (v-x)c + (y-u)h)^2)]^{3/2}} $$</p> <p>where $(u,y,x,v,c,h) \in \mathbb{R}^5 \times \mathbb{R}^+$. Although the integral is uglier, it satisfyingly has no absolute value operator.</p> Answer 2: <p>If anyone is still interested, Antonio Lerario and I recently published a paper: <a href="https://link.springer.com/article/10.1007/s40598-020-00160-w" rel="nofollow noreferrer">Probabilistic Schubert Calculus: Asymptotics</a>, in which we give a more convenient formula to compute this number.</p> <p>What was called before <span class="math-container">$\operatorname{edeg}G(2,4)$</span> is now denoted by <span class="math-container">$\delta_{1,3}$</span> and we give a line integral formula in Proposition 24:</p> <p><span class="math-container">\begin{aligned} \delta _{1,3}=-6 \pi ^{4}\int _{0}^1 L(u)^{2}\mathrm {sinh}(w(u))w'(u)\mathrm{d}u \end{aligned}</span> where <span class="math-container">$L=F⋅G$</span> and <span class="math-container">$w=log(F/G)$</span> with <span class="math-container">\begin{aligned} F(u)&amp;:=\int _{0}^{\pi /2} \frac{u \ \sin ^2(\theta )}{\sqrt{\cos ^2\theta +u^2 \sin ^2\theta }}\mathrm {d}\theta \\ G (u)&amp;:=\int _{0}^{\pi /2} \frac{\sin ^2(\theta )}{\sqrt{\sin ^2\theta +u^2 \cos ^2\theta }}\mathrm {d}\theta . \end{aligned}</span></p> <p>We didn't run any advanced numerical evaluation but this looks much nicer than the previous formula. I hope this will help. It is still unknown to us if this number can be expressed as a closed formula using special functions.</p> Answer 3: <p>The integrand is periodic modulo $\pi$ in each variable, so it suffices to integrate each variable over $[0, \pi]$ and replace the constant factor by $2^{-7}$.</p> <p>If we were to apply a change of variables (e.g. set $x = \cos(s_1)$ and similarly for the other five variables), we would have an integral of a piecewise-algebraic function which thus belongs to Kontsevich and Zagier's 'ring of periods'. Alas, there is no proven algorithm for determining whether periods are expressible in terms of elementary functions.</p> <p>Now, using Gauss-Legendre integration with $N$ points in each variable, we can approximate it with a weighted sum of $N^6$ evaluations of the integrand. Since the absolute value of the determinant is unchanged under permuting the columns, and zero whenever two columns are equal, we can reduce this to $\binom{N^2}{3}$ function evaluations.</p> <p>I managed to implement this in C code, where each iteration, amortized, only takes four floating-point additions, three multiplications, and an absolute-value calculation. With several hours running on a 144-core machine, it could do the calculation for both $N = 401$ and $N = 409$. The results were 1.726230867 and 1.726230885, respectively.</p> <p>Consequently, the integral is roughly 1.726230876, with an expected error on the order of $10^{-8}$.</p> <hr> <p><strong>EDIT:</strong> Matt F. posted an algebraic form for the integral, by taking $u = \cot(t1)$, $v = \cot(s2)$, $w = \cot(t3)$, $x = \cot(s1)$, $y = \cot(t2)$, $z = \cot(s3)$:</p> <p>$$ I = 2^{-7} \int \dfrac{|uv-vw+wx-xy+yz-zu|}{[(1+u^2)(1+v^2)(1+w^2)(1+x^2)(1+y^2)(1+z^2)]^{3/2}} $$</p> <p>where the integral is taken over $(u,v,w,x,y,z) \in \mathbb{R}^6$. Observing that the numerator is twice the area of a triangle with vertices $(u, x), (y, v), (w, z)$, we can perform a change of variables by setting:</p> <p>$$ w = u + (y-u)c - (v-x)h \textrm{ ; } z = x + (v-x)c + (y-u)h $$</p> <p>where $c \in \mathbb{R}$ and $h \in \mathbb{R}^{+}$. Taking the Jacobian into account, the integral becomes:</p> <p>$$ I = 2^{-6} \int \dfrac{h[(y-u)^2+(v-x)^2]^2}{[(1+u^2)(1+y^2)(1+(u + (y-u)c - (v-x)h)^2)(1+x^2)(1+v^2)(1+(x + (v-x)c + (y-u)h)^2)]^{3/2}} $$</p> <p>where $(u,y,x,v,c,h) \in \mathbb{R}^5 \times \mathbb{R}^+$. Although the integral is uglier, it satisfyingly has no absolute value operator.</p> Answer 4: <p>If anyone is still interested, Antonio Lerario and I recently published a paper: <a href="https://link.springer.com/article/10.1007/s40598-020-00160-w" rel="nofollow noreferrer">Probabilistic Schubert Calculus: Asymptotics</a>, in which we give a more convenient formula to compute this number.</p> <p>What was called before <span class="math-container">$\operatorname{edeg}G(2,4)$</span> is now denoted by <span class="math-container">$\delta_{1,3}$</span> and we give a line integral formula in Proposition 24:</p> <p><span class="math-container">\begin{aligned} \delta _{1,3}=-6 \pi ^{4}\int _{0}^1 L(u)^{2}\mathrm {sinh}(w(u))w'(u)\mathrm{d}u \end{aligned}</span> where <span class="math-container">$L=F⋅G$</span> and <span class="math-container">$w=log(F/G)$</span> with <span class="math-container">\begin{aligned} F(u)&amp;:=\int _{0}^{\pi /2} \frac{u \ \sin ^2(\theta )}{\sqrt{\cos ^2\theta +u^2 \sin ^2\theta }}\mathrm {d}\theta \\ G (u)&amp;:=\int _{0}^{\pi /2} \frac{\sin ^2(\theta )}{\sqrt{\sin ^2\theta +u^2 \cos ^2\theta }}\mathrm {d}\theta . \end{aligned}</span></p> <p>We didn't run any advanced numerical evaluation but this looks much nicer than the previous formula. I hope this will help. It is still unknown to us if this number can be expressed as a closed formula using special functions.</p> Answer 5: <p>The integrand is periodic modulo $\pi$ in each variable, so it suffices to integrate each variable over $[0, \pi]$ and replace the constant factor by $2^{-7}$.</p> <p>If we were to apply a change of variables (e.g. set $x = \cos(s_1)$ and similarly for the other five variables), we would have an integral of a piecewise-algebraic function which thus belongs to Kontsevich and Zagier's 'ring of periods'. Alas, there is no proven algorithm for determining whether periods are expressible in terms of elementary functions.</p> <p>Now, using Gauss-Legendre integration with $N$ points in each variable, we can approximate it with a weighted sum of $N^6$ evaluations of the integrand. Since the absolute value of the determinant is unchanged under permuting the columns, and zero whenever two columns are equal, we can reduce this to $\binom{N^2}{3}$ function evaluations.</p> <p>I managed to implement this in C code, where each iteration, amortized, only takes four floating-point additions, three multiplications, and an absolute-value calculation. With several hours running on a 144-core machine, it could do the calculation for both $N = 401$ and $N = 409$. The results were 1.726230867 and 1.726230885, respectively.</p> <p>Consequently, the integral is roughly 1.726230876, with an expected error on the order of $10^{-8}$.</p> <hr> <p><strong>EDIT:</strong> Matt F. posted an algebraic form for the integral, by taking $u = \cot(t1)$, $v = \cot(s2)$, $w = \cot(t3)$, $x = \cot(s1)$, $y = \cot(t2)$, $z = \cot(s3)$:</p> <p>$$ I = 2^{-7} \int \dfrac{|uv-vw+wx-xy+yz-zu|}{[(1+u^2)(1+v^2)(1+w^2)(1+x^2)(1+y^2)(1+z^2)]^{3/2}} $$</p> <p>where the integral is taken over $(u,v,w,x,y,z) \in \mathbb{R}^6$. Observing that the numerator is twice the area of a triangle with vertices $(u, x), (y, v), (w, z)$, we can perform a change of variables by setting:</p> <p>$$ w = u + (y-u)c - (v-x)h \textrm{ ; } z = x + (v-x)c + (y-u)h $$</p> <p>where $c \in \mathbb{R}$ and $h \in \mathbb{R}^{+}$. Taking the Jacobian into account, the integral becomes:</p> <p>$$ I = 2^{-6} \int \dfrac{h[(y-u)^2+(v-x)^2]^2}{[(1+u^2)(1+y^2)(1+(u + (y-u)c - (v-x)h)^2)(1+x^2)(1+v^2)(1+(x + (v-x)c + (y-u)h)^2)]^{3/2}} $$</p> <p>where $(u,y,x,v,c,h) \in \mathbb{R}^5 \times \mathbb{R}^+$. Although the integral is uglier, it satisfyingly has no absolute value operator.</p> Answer 6: <p>If anyone is still interested, Antonio Lerario and I recently published a paper: <a href="https://link.springer.com/article/10.1007/s40598-020-00160-w" rel="nofollow noreferrer">Probabilistic Schubert Calculus: Asymptotics</a>, in which we give a more convenient formula to compute this number.</p> <p>What was called before <span class="math-container">$\operatorname{edeg}G(2,4)$</span> is now denoted by <span class="math-container">$\delta_{1,3}$</span> and we give a line integral formula in Proposition 24:</p> <p><span class="math-container">\begin{aligned} \delta _{1,3}=-6 \pi ^{4}\int _{0}^1 L(u)^{2}\mathrm {sinh}(w(u))w'(u)\mathrm{d}u \end{aligned}</span> where <span class="math-container">$L=F⋅G$</span> and <span class="math-container">$w=log(F/G)$</span> with <span class="math-container">\begin{aligned} F(u)&amp;:=\int _{0}^{\pi /2} \frac{u \ \sin ^2(\theta )}{\sqrt{\cos ^2\theta +u^2 \sin ^2\theta }}\mathrm {d}\theta \\ G (u)&amp;:=\int _{0}^{\pi /2} \frac{\sin ^2(\theta )}{\sqrt{\sin ^2\theta +u^2 \cos ^2\theta }}\mathrm {d}\theta . \end{aligned}</span></p> <p>We didn't run any advanced numerical evaluation but this looks much nicer than the previous formula. I hope this will help. It is still unknown to us if this number can be expressed as a closed formula using special functions.</p>
Radon-Nikodym theorem for non-sigma finite measures
258,936
13
https://mathoverflow.net/questions/258936/radon-nikodym-theorem-for-non-sigma-finite-measures
<p>Let $(X,\mathcal M, \mu)$ be a measured space where $\mu$ is a positive measure. Let $\lambda$ be a complex measure on $(X,\mathcal M)$. When $\mu$ is sigma-finite, the Radon-Nikodym theorem provides a decomposition of $\lambda$ in a sum of an absolutely continuous measure wrt $\mu$ plus a singular measure wrt $\mu$.</p> <p>Question. Is that statement still true without the sigma-finiteness of $\mu$? It does seems to be so easy for the case of the Hausdorff measures of dimension $d&lt;n$ in $\mathbb R^n$.</p>
[ { "answer_id": 258998, "score": 15, "text": "<p>The measure spaces for which all of the nice theorems about $\\sigma$-finite measures (except those about product measures, which require other generalizations) generalize are called localizable measure spaces. There are a number of equivalent definitions of localizable measure spaces; perhaps the simplest states that a positive measure $\\mu$ is <em>localizable</em> if it is semifinite and has essential suprema, and a complex measure is localizable if its total variation is.</p>\n\n<p>A measure is <em>semifinite</em> if whenever $\\mu(E) = \\infty$ there exists a measurable $F \\subseteq E$ such that $0 &lt; \\mu(F) &lt; \\infty$. A measure has <em>essential suprema</em> if for every family $\\mathcal{E}$ of measurable sets there exists a measurable set $H$ such that</p>\n\n<ul>\n<li>$E \\setminus H$ is null for every $E \\in \\mathcal{E}$, and</li>\n<li>if $G$ is measurable and $E \\setminus G$ is negligible for every $E \\in \\mathcal{E}$, then $H \\setminus G$ is negligible.</li>\n</ul>\n\n<p>The most commonly used equivalent definition of a localizable measure space is that real-valued $L^\\infty(\\mu)$ is Dedekind-complete as a lattice. Essentially all measures that arise in applications are localizable, e.g. all regular Borel measures on locally compact Hausdorff spaces are localizable.</p>\n\n<p>The theorem that you would like to generalize is normally called the Lebesgue Decomposition Theorem rather than the Radon-Nikodym Theorem, which is about the representation of absolutely continuous measures in terms of integration against functions. You are missing the hypothesis that $\\lambda$ is also $\\sigma$-finite, or at least that $\\mu + \\lambda$ is $\\sigma$-finite.</p>\n\n<p>If $\\mu$ and $\\nu$ are measures on the same measurable space such that $\\mu + \\nu$ is localizable, then there exists a decomposition $\\mu = \\mu_c + \\mu_s$ where $\\mu_c$ is absolutely continuous with respect to $\\nu$ and $\\mu_s$ is singular. For a simple proof, see <a href=\"http://link.springer.com/article/10.1007/BF02052840\" rel=\"noreferrer\">Decomposition and Representation Theorems in Measure Theory</a> by Kelley.</p>\n" }, { "answer_id": 259008, "score": 7, "text": "<p>One can formulate the Radon-Nikodym so that it applies to all measure spaces without any restriction. Let $(X,\\Sigma,\\nu)$ be a measure space and $\\nu$ be a real valued signed measure on $\\Sigma$. </p>\n\n<p>Then $\\nu$ is absolutely continuous with respect to $\\mu$ if and only if for all $\\epsilon&gt;0$ there exists $\\delta&gt;0$ such that for all $A\\in\\Sigma$, $\\mu(A)\\leq\\delta$ implies $|\\nu(A)|\\leq\\epsilon$.</p>\n\n<p>We say that $\\nu$ is <em>truly continuous</em> with respect to $\\mu$ if and only for all $\\epsilon&gt;0$ there exists $\\delta&gt;0$ and $F\\in\\Sigma$ with $\\mu(F)&lt;\\infty$ such that for all $A\\in\\Sigma$, $\\mu(A\\cap F)\\leq\\delta$ implies $|\\nu(A)|\\leq\\epsilon$.</p>\n\n<p>If $(X,\\Sigma,\\nu)$ is $\\sigma$-finite, then a real-valued signed measure $\\nu$ on $\\Sigma$ is truly continuous with respect of $\\mu$ if and only if it is absolutely continuous with respect to $\\mu$.</p>\n\n<p><strong>Theorem:</strong> Let $(X,\\Sigma,\\nu)$ be a measure space and $\\nu$ be a real valued signed measure on $\\Sigma$. Then $\\nu$ is truly continuous with respect to $\\mu$ if and only if there exists a measurable function $f:\\Sigma\\to \\mathbb{R}$ such that $$\\nu(A)=\\int_A f~\\text{d}\\mu$$ for all $A\\in\\Sigma$.</p>\n\n<p>This formulation of the Radon-Nikodym is from the second volume of David Fremlin's <a href=\"https://www.essex.ac.uk/maths/people/fremlin/mt.htm\" rel=\"noreferrer\">treatise on measure theory</a>. Fremlin doesn't assume $\\nu$ to be countably additive, which requires additional distinctions. There is also a treatment in Section 5.4 of the book <a href=\"http://www.ems-ph.org/books/book.php?proj_nr=201\" rel=\"noreferrer\">Measure and Integration</a> by Dietmar Salamon (preprint available <a href=\"https://people.math.ethz.ch/~salamon/PREPRINTS/measure.pdf\" rel=\"noreferrer\">here</a>), which also contains a useful characterization by Heinz König of being truly continuous in terms of absolute continuity and an inner regularity property</p>\n" }, { "answer_id": 258998, "score": 15, "text": "<p>The measure spaces for which all of the nice theorems about $\\sigma$-finite measures (except those about product measures, which require other generalizations) generalize are called localizable measure spaces. There are a number of equivalent definitions of localizable measure spaces; perhaps the simplest states that a positive measure $\\mu$ is <em>localizable</em> if it is semifinite and has essential suprema, and a complex measure is localizable if its total variation is.</p>\n\n<p>A measure is <em>semifinite</em> if whenever $\\mu(E) = \\infty$ there exists a measurable $F \\subseteq E$ such that $0 &lt; \\mu(F) &lt; \\infty$. A measure has <em>essential suprema</em> if for every family $\\mathcal{E}$ of measurable sets there exists a measurable set $H$ such that</p>\n\n<ul>\n<li>$E \\setminus H$ is null for every $E \\in \\mathcal{E}$, and</li>\n<li>if $G$ is measurable and $E \\setminus G$ is negligible for every $E \\in \\mathcal{E}$, then $H \\setminus G$ is negligible.</li>\n</ul>\n\n<p>The most commonly used equivalent definition of a localizable measure space is that real-valued $L^\\infty(\\mu)$ is Dedekind-complete as a lattice. Essentially all measures that arise in applications are localizable, e.g. all regular Borel measures on locally compact Hausdorff spaces are localizable.</p>\n\n<p>The theorem that you would like to generalize is normally called the Lebesgue Decomposition Theorem rather than the Radon-Nikodym Theorem, which is about the representation of absolutely continuous measures in terms of integration against functions. You are missing the hypothesis that $\\lambda$ is also $\\sigma$-finite, or at least that $\\mu + \\lambda$ is $\\sigma$-finite.</p>\n\n<p>If $\\mu$ and $\\nu$ are measures on the same measurable space such that $\\mu + \\nu$ is localizable, then there exists a decomposition $\\mu = \\mu_c + \\mu_s$ where $\\mu_c$ is absolutely continuous with respect to $\\nu$ and $\\mu_s$ is singular. For a simple proof, see <a href=\"http://link.springer.com/article/10.1007/BF02052840\" rel=\"noreferrer\">Decomposition and Representation Theorems in Measure Theory</a> by Kelley.</p>\n" }, { "answer_id": 259008, "score": 7, "text": "<p>One can formulate the Radon-Nikodym so that it applies to all measure spaces without any restriction. Let $(X,\\Sigma,\\nu)$ be a measure space and $\\nu$ be a real valued signed measure on $\\Sigma$. </p>\n\n<p>Then $\\nu$ is absolutely continuous with respect to $\\mu$ if and only if for all $\\epsilon&gt;0$ there exists $\\delta&gt;0$ such that for all $A\\in\\Sigma$, $\\mu(A)\\leq\\delta$ implies $|\\nu(A)|\\leq\\epsilon$.</p>\n\n<p>We say that $\\nu$ is <em>truly continuous</em> with respect to $\\mu$ if and only for all $\\epsilon&gt;0$ there exists $\\delta&gt;0$ and $F\\in\\Sigma$ with $\\mu(F)&lt;\\infty$ such that for all $A\\in\\Sigma$, $\\mu(A\\cap F)\\leq\\delta$ implies $|\\nu(A)|\\leq\\epsilon$.</p>\n\n<p>If $(X,\\Sigma,\\nu)$ is $\\sigma$-finite, then a real-valued signed measure $\\nu$ on $\\Sigma$ is truly continuous with respect of $\\mu$ if and only if it is absolutely continuous with respect to $\\mu$.</p>\n\n<p><strong>Theorem:</strong> Let $(X,\\Sigma,\\nu)$ be a measure space and $\\nu$ be a real valued signed measure on $\\Sigma$. Then $\\nu$ is truly continuous with respect to $\\mu$ if and only if there exists a measurable function $f:\\Sigma\\to \\mathbb{R}$ such that $$\\nu(A)=\\int_A f~\\text{d}\\mu$$ for all $A\\in\\Sigma$.</p>\n\n<p>This formulation of the Radon-Nikodym is from the second volume of David Fremlin's <a href=\"https://www.essex.ac.uk/maths/people/fremlin/mt.htm\" rel=\"noreferrer\">treatise on measure theory</a>. Fremlin doesn't assume $\\nu$ to be countably additive, which requires additional distinctions. There is also a treatment in Section 5.4 of the book <a href=\"http://www.ems-ph.org/books/book.php?proj_nr=201\" rel=\"noreferrer\">Measure and Integration</a> by Dietmar Salamon (preprint available <a href=\"https://people.math.ethz.ch/~salamon/PREPRINTS/measure.pdf\" rel=\"noreferrer\">here</a>), which also contains a useful characterization by Heinz König of being truly continuous in terms of absolute continuity and an inner regularity property</p>\n" }, { "answer_id": 258998, "score": 15, "text": "<p>The measure spaces for which all of the nice theorems about $\\sigma$-finite measures (except those about product measures, which require other generalizations) generalize are called localizable measure spaces. There are a number of equivalent definitions of localizable measure spaces; perhaps the simplest states that a positive measure $\\mu$ is <em>localizable</em> if it is semifinite and has essential suprema, and a complex measure is localizable if its total variation is.</p>\n\n<p>A measure is <em>semifinite</em> if whenever $\\mu(E) = \\infty$ there exists a measurable $F \\subseteq E$ such that $0 &lt; \\mu(F) &lt; \\infty$. A measure has <em>essential suprema</em> if for every family $\\mathcal{E}$ of measurable sets there exists a measurable set $H$ such that</p>\n\n<ul>\n<li>$E \\setminus H$ is null for every $E \\in \\mathcal{E}$, and</li>\n<li>if $G$ is measurable and $E \\setminus G$ is negligible for every $E \\in \\mathcal{E}$, then $H \\setminus G$ is negligible.</li>\n</ul>\n\n<p>The most commonly used equivalent definition of a localizable measure space is that real-valued $L^\\infty(\\mu)$ is Dedekind-complete as a lattice. Essentially all measures that arise in applications are localizable, e.g. all regular Borel measures on locally compact Hausdorff spaces are localizable.</p>\n\n<p>The theorem that you would like to generalize is normally called the Lebesgue Decomposition Theorem rather than the Radon-Nikodym Theorem, which is about the representation of absolutely continuous measures in terms of integration against functions. You are missing the hypothesis that $\\lambda$ is also $\\sigma$-finite, or at least that $\\mu + \\lambda$ is $\\sigma$-finite.</p>\n\n<p>If $\\mu$ and $\\nu$ are measures on the same measurable space such that $\\mu + \\nu$ is localizable, then there exists a decomposition $\\mu = \\mu_c + \\mu_s$ where $\\mu_c$ is absolutely continuous with respect to $\\nu$ and $\\mu_s$ is singular. For a simple proof, see <a href=\"http://link.springer.com/article/10.1007/BF02052840\" rel=\"noreferrer\">Decomposition and Representation Theorems in Measure Theory</a> by Kelley.</p>\n" }, { "answer_id": 259008, "score": 7, "text": "<p>One can formulate the Radon-Nikodym so that it applies to all measure spaces without any restriction. Let $(X,\\Sigma,\\nu)$ be a measure space and $\\nu$ be a real valued signed measure on $\\Sigma$. </p>\n\n<p>Then $\\nu$ is absolutely continuous with respect to $\\mu$ if and only if for all $\\epsilon&gt;0$ there exists $\\delta&gt;0$ such that for all $A\\in\\Sigma$, $\\mu(A)\\leq\\delta$ implies $|\\nu(A)|\\leq\\epsilon$.</p>\n\n<p>We say that $\\nu$ is <em>truly continuous</em> with respect to $\\mu$ if and only for all $\\epsilon&gt;0$ there exists $\\delta&gt;0$ and $F\\in\\Sigma$ with $\\mu(F)&lt;\\infty$ such that for all $A\\in\\Sigma$, $\\mu(A\\cap F)\\leq\\delta$ implies $|\\nu(A)|\\leq\\epsilon$.</p>\n\n<p>If $(X,\\Sigma,\\nu)$ is $\\sigma$-finite, then a real-valued signed measure $\\nu$ on $\\Sigma$ is truly continuous with respect of $\\mu$ if and only if it is absolutely continuous with respect to $\\mu$.</p>\n\n<p><strong>Theorem:</strong> Let $(X,\\Sigma,\\nu)$ be a measure space and $\\nu$ be a real valued signed measure on $\\Sigma$. Then $\\nu$ is truly continuous with respect to $\\mu$ if and only if there exists a measurable function $f:\\Sigma\\to \\mathbb{R}$ such that $$\\nu(A)=\\int_A f~\\text{d}\\mu$$ for all $A\\in\\Sigma$.</p>\n\n<p>This formulation of the Radon-Nikodym is from the second volume of David Fremlin's <a href=\"https://www.essex.ac.uk/maths/people/fremlin/mt.htm\" rel=\"noreferrer\">treatise on measure theory</a>. Fremlin doesn't assume $\\nu$ to be countably additive, which requires additional distinctions. There is also a treatment in Section 5.4 of the book <a href=\"http://www.ems-ph.org/books/book.php?proj_nr=201\" rel=\"noreferrer\">Measure and Integration</a> by Dietmar Salamon (preprint available <a href=\"https://people.math.ethz.ch/~salamon/PREPRINTS/measure.pdf\" rel=\"noreferrer\">here</a>), which also contains a useful characterization by Heinz König of being truly continuous in terms of absolute continuity and an inner regularity property</p>\n" } ]
Question: Radon-Nikodym theorem for non-sigma finite measures Answer 1: <p>The measure spaces for which all of the nice theorems about $\sigma$-finite measures (except those about product measures, which require other generalizations) generalize are called localizable measure spaces. There are a number of equivalent definitions of localizable measure spaces; perhaps the simplest states that a positive measure $\mu$ is <em>localizable</em> if it is semifinite and has essential suprema, and a complex measure is localizable if its total variation is.</p> <p>A measure is <em>semifinite</em> if whenever $\mu(E) = \infty$ there exists a measurable $F \subseteq E$ such that $0 &lt; \mu(F) &lt; \infty$. A measure has <em>essential suprema</em> if for every family $\mathcal{E}$ of measurable sets there exists a measurable set $H$ such that</p> <ul> <li>$E \setminus H$ is null for every $E \in \mathcal{E}$, and</li> <li>if $G$ is measurable and $E \setminus G$ is negligible for every $E \in \mathcal{E}$, then $H \setminus G$ is negligible.</li> </ul> <p>The most commonly used equivalent definition of a localizable measure space is that real-valued $L^\infty(\mu)$ is Dedekind-complete as a lattice. Essentially all measures that arise in applications are localizable, e.g. all regular Borel measures on locally compact Hausdorff spaces are localizable.</p> <p>The theorem that you would like to generalize is normally called the Lebesgue Decomposition Theorem rather than the Radon-Nikodym Theorem, which is about the representation of absolutely continuous measures in terms of integration against functions. You are missing the hypothesis that $\lambda$ is also $\sigma$-finite, or at least that $\mu + \lambda$ is $\sigma$-finite.</p> <p>If $\mu$ and $\nu$ are measures on the same measurable space such that $\mu + \nu$ is localizable, then there exists a decomposition $\mu = \mu_c + \mu_s$ where $\mu_c$ is absolutely continuous with respect to $\nu$ and $\mu_s$ is singular. For a simple proof, see <a href="http://link.springer.com/article/10.1007/BF02052840" rel="noreferrer">Decomposition and Representation Theorems in Measure Theory</a> by Kelley.</p> Answer 2: <p>One can formulate the Radon-Nikodym so that it applies to all measure spaces without any restriction. Let $(X,\Sigma,\nu)$ be a measure space and $\nu$ be a real valued signed measure on $\Sigma$. </p> <p>Then $\nu$ is absolutely continuous with respect to $\mu$ if and only if for all $\epsilon&gt;0$ there exists $\delta&gt;0$ such that for all $A\in\Sigma$, $\mu(A)\leq\delta$ implies $|\nu(A)|\leq\epsilon$.</p> <p>We say that $\nu$ is <em>truly continuous</em> with respect to $\mu$ if and only for all $\epsilon&gt;0$ there exists $\delta&gt;0$ and $F\in\Sigma$ with $\mu(F)&lt;\infty$ such that for all $A\in\Sigma$, $\mu(A\cap F)\leq\delta$ implies $|\nu(A)|\leq\epsilon$.</p> <p>If $(X,\Sigma,\nu)$ is $\sigma$-finite, then a real-valued signed measure $\nu$ on $\Sigma$ is truly continuous with respect of $\mu$ if and only if it is absolutely continuous with respect to $\mu$.</p> <p><strong>Theorem:</strong> Let $(X,\Sigma,\nu)$ be a measure space and $\nu$ be a real valued signed measure on $\Sigma$. Then $\nu$ is truly continuous with respect to $\mu$ if and only if there exists a measurable function $f:\Sigma\to \mathbb{R}$ such that $$\nu(A)=\int_A f~\text{d}\mu$$ for all $A\in\Sigma$.</p> <p>This formulation of the Radon-Nikodym is from the second volume of David Fremlin's <a href="https://www.essex.ac.uk/maths/people/fremlin/mt.htm" rel="noreferrer">treatise on measure theory</a>. Fremlin doesn't assume $\nu$ to be countably additive, which requires additional distinctions. There is also a treatment in Section 5.4 of the book <a href="http://www.ems-ph.org/books/book.php?proj_nr=201" rel="noreferrer">Measure and Integration</a> by Dietmar Salamon (preprint available <a href="https://people.math.ethz.ch/~salamon/PREPRINTS/measure.pdf" rel="noreferrer">here</a>), which also contains a useful characterization by Heinz König of being truly continuous in terms of absolute continuity and an inner regularity property</p> Answer 3: <p>The measure spaces for which all of the nice theorems about $\sigma$-finite measures (except those about product measures, which require other generalizations) generalize are called localizable measure spaces. There are a number of equivalent definitions of localizable measure spaces; perhaps the simplest states that a positive measure $\mu$ is <em>localizable</em> if it is semifinite and has essential suprema, and a complex measure is localizable if its total variation is.</p> <p>A measure is <em>semifinite</em> if whenever $\mu(E) = \infty$ there exists a measurable $F \subseteq E$ such that $0 &lt; \mu(F) &lt; \infty$. A measure has <em>essential suprema</em> if for every family $\mathcal{E}$ of measurable sets there exists a measurable set $H$ such that</p> <ul> <li>$E \setminus H$ is null for every $E \in \mathcal{E}$, and</li> <li>if $G$ is measurable and $E \setminus G$ is negligible for every $E \in \mathcal{E}$, then $H \setminus G$ is negligible.</li> </ul> <p>The most commonly used equivalent definition of a localizable measure space is that real-valued $L^\infty(\mu)$ is Dedekind-complete as a lattice. Essentially all measures that arise in applications are localizable, e.g. all regular Borel measures on locally compact Hausdorff spaces are localizable.</p> <p>The theorem that you would like to generalize is normally called the Lebesgue Decomposition Theorem rather than the Radon-Nikodym Theorem, which is about the representation of absolutely continuous measures in terms of integration against functions. You are missing the hypothesis that $\lambda$ is also $\sigma$-finite, or at least that $\mu + \lambda$ is $\sigma$-finite.</p> <p>If $\mu$ and $\nu$ are measures on the same measurable space such that $\mu + \nu$ is localizable, then there exists a decomposition $\mu = \mu_c + \mu_s$ where $\mu_c$ is absolutely continuous with respect to $\nu$ and $\mu_s$ is singular. For a simple proof, see <a href="http://link.springer.com/article/10.1007/BF02052840" rel="noreferrer">Decomposition and Representation Theorems in Measure Theory</a> by Kelley.</p> Answer 4: <p>One can formulate the Radon-Nikodym so that it applies to all measure spaces without any restriction. Let $(X,\Sigma,\nu)$ be a measure space and $\nu$ be a real valued signed measure on $\Sigma$. </p> <p>Then $\nu$ is absolutely continuous with respect to $\mu$ if and only if for all $\epsilon&gt;0$ there exists $\delta&gt;0$ such that for all $A\in\Sigma$, $\mu(A)\leq\delta$ implies $|\nu(A)|\leq\epsilon$.</p> <p>We say that $\nu$ is <em>truly continuous</em> with respect to $\mu$ if and only for all $\epsilon&gt;0$ there exists $\delta&gt;0$ and $F\in\Sigma$ with $\mu(F)&lt;\infty$ such that for all $A\in\Sigma$, $\mu(A\cap F)\leq\delta$ implies $|\nu(A)|\leq\epsilon$.</p> <p>If $(X,\Sigma,\nu)$ is $\sigma$-finite, then a real-valued signed measure $\nu$ on $\Sigma$ is truly continuous with respect of $\mu$ if and only if it is absolutely continuous with respect to $\mu$.</p> <p><strong>Theorem:</strong> Let $(X,\Sigma,\nu)$ be a measure space and $\nu$ be a real valued signed measure on $\Sigma$. Then $\nu$ is truly continuous with respect to $\mu$ if and only if there exists a measurable function $f:\Sigma\to \mathbb{R}$ such that $$\nu(A)=\int_A f~\text{d}\mu$$ for all $A\in\Sigma$.</p> <p>This formulation of the Radon-Nikodym is from the second volume of David Fremlin's <a href="https://www.essex.ac.uk/maths/people/fremlin/mt.htm" rel="noreferrer">treatise on measure theory</a>. Fremlin doesn't assume $\nu$ to be countably additive, which requires additional distinctions. There is also a treatment in Section 5.4 of the book <a href="http://www.ems-ph.org/books/book.php?proj_nr=201" rel="noreferrer">Measure and Integration</a> by Dietmar Salamon (preprint available <a href="https://people.math.ethz.ch/~salamon/PREPRINTS/measure.pdf" rel="noreferrer">here</a>), which also contains a useful characterization by Heinz König of being truly continuous in terms of absolute continuity and an inner regularity property</p> Answer 5: <p>The measure spaces for which all of the nice theorems about $\sigma$-finite measures (except those about product measures, which require other generalizations) generalize are called localizable measure spaces. There are a number of equivalent definitions of localizable measure spaces; perhaps the simplest states that a positive measure $\mu$ is <em>localizable</em> if it is semifinite and has essential suprema, and a complex measure is localizable if its total variation is.</p> <p>A measure is <em>semifinite</em> if whenever $\mu(E) = \infty$ there exists a measurable $F \subseteq E$ such that $0 &lt; \mu(F) &lt; \infty$. A measure has <em>essential suprema</em> if for every family $\mathcal{E}$ of measurable sets there exists a measurable set $H$ such that</p> <ul> <li>$E \setminus H$ is null for every $E \in \mathcal{E}$, and</li> <li>if $G$ is measurable and $E \setminus G$ is negligible for every $E \in \mathcal{E}$, then $H \setminus G$ is negligible.</li> </ul> <p>The most commonly used equivalent definition of a localizable measure space is that real-valued $L^\infty(\mu)$ is Dedekind-complete as a lattice. Essentially all measures that arise in applications are localizable, e.g. all regular Borel measures on locally compact Hausdorff spaces are localizable.</p> <p>The theorem that you would like to generalize is normally called the Lebesgue Decomposition Theorem rather than the Radon-Nikodym Theorem, which is about the representation of absolutely continuous measures in terms of integration against functions. You are missing the hypothesis that $\lambda$ is also $\sigma$-finite, or at least that $\mu + \lambda$ is $\sigma$-finite.</p> <p>If $\mu$ and $\nu$ are measures on the same measurable space such that $\mu + \nu$ is localizable, then there exists a decomposition $\mu = \mu_c + \mu_s$ where $\mu_c$ is absolutely continuous with respect to $\nu$ and $\mu_s$ is singular. For a simple proof, see <a href="http://link.springer.com/article/10.1007/BF02052840" rel="noreferrer">Decomposition and Representation Theorems in Measure Theory</a> by Kelley.</p> Answer 6: <p>One can formulate the Radon-Nikodym so that it applies to all measure spaces without any restriction. Let $(X,\Sigma,\nu)$ be a measure space and $\nu$ be a real valued signed measure on $\Sigma$. </p> <p>Then $\nu$ is absolutely continuous with respect to $\mu$ if and only if for all $\epsilon&gt;0$ there exists $\delta&gt;0$ such that for all $A\in\Sigma$, $\mu(A)\leq\delta$ implies $|\nu(A)|\leq\epsilon$.</p> <p>We say that $\nu$ is <em>truly continuous</em> with respect to $\mu$ if and only for all $\epsilon&gt;0$ there exists $\delta&gt;0$ and $F\in\Sigma$ with $\mu(F)&lt;\infty$ such that for all $A\in\Sigma$, $\mu(A\cap F)\leq\delta$ implies $|\nu(A)|\leq\epsilon$.</p> <p>If $(X,\Sigma,\nu)$ is $\sigma$-finite, then a real-valued signed measure $\nu$ on $\Sigma$ is truly continuous with respect of $\mu$ if and only if it is absolutely continuous with respect to $\mu$.</p> <p><strong>Theorem:</strong> Let $(X,\Sigma,\nu)$ be a measure space and $\nu$ be a real valued signed measure on $\Sigma$. Then $\nu$ is truly continuous with respect to $\mu$ if and only if there exists a measurable function $f:\Sigma\to \mathbb{R}$ such that $$\nu(A)=\int_A f~\text{d}\mu$$ for all $A\in\Sigma$.</p> <p>This formulation of the Radon-Nikodym is from the second volume of David Fremlin's <a href="https://www.essex.ac.uk/maths/people/fremlin/mt.htm" rel="noreferrer">treatise on measure theory</a>. Fremlin doesn't assume $\nu$ to be countably additive, which requires additional distinctions. There is also a treatment in Section 5.4 of the book <a href="http://www.ems-ph.org/books/book.php?proj_nr=201" rel="noreferrer">Measure and Integration</a> by Dietmar Salamon (preprint available <a href="https://people.math.ethz.ch/~salamon/PREPRINTS/measure.pdf" rel="noreferrer">here</a>), which also contains a useful characterization by Heinz König of being truly continuous in terms of absolute continuity and an inner regularity property</p>
Memorable ordinals
259,100
13
https://mathoverflow.net/questions/259100/memorable-ordinals
<p>Suppose $V=L$ + reasonable hypotheses (e.g. "ZFC has a countable transitive model"). Call a countable ordinal $\alpha$ <em>memorable</em> if for some countable $\beta$, $\alpha$ is definable <em>without parameters</em> in every $L_\gamma$ with $\beta&lt;\gamma&lt;\omega_1$.</p> <p>My question is:</p> <blockquote> <p>Are there uncountably many memorable ordinals?</p> </blockquote> <hr> <p>Some comments, right off the bat:</p> <ul> <li><p>The answer is trivially "no" for <em>uniformly</em> memorable ordinals, that is, ordinals which are coboundedly definable without parameters <em>by the same formula</em>. However, since the defining formula is allowed to vary with $\gamma$, this doesn't work.</p></li> <li><p>Towards a positive answer, note that there are uncountably many countable $\theta$ such that $L_\theta$ is pointwise definable (call such an ordinal "insightful"); this was proved by <a href="https://arxiv.org/pdf/1105.4597v2.pdf" rel="noreferrer">Hamkins, Linetsky, and Reitz</a>. However, this doesn't actually resolve the issue: for a fixed $\alpha$, there may be many countable $\eta&gt;\alpha$ with no insightful ordinals $&gt;\alpha$ which are definable without parameters in $L_\eta$ (e.g. if there is no greatest insightful ordinal $&lt;\eta$, this seems a distinct possibility).</p></li> </ul> <hr> <p>Interestingly, this question has possibly interesting variants even if $L$ is a very tiny subclass of $V$! Given any hierarchy $(M_\alpha)_{\alpha\in\omega_1}$ with $M_\alpha\cap ON=\alpha$, we can ask whether uncountably many $\alpha$ satisfy "For all sufficiently large $\beta&lt;\omega_1$, $\alpha$ is parameter-free definable in $M_\beta$." Now, we can trivially construct examples where the answer is "yes": namely, let $M_\alpha$ be $L_{\eta_\alpha}$ where $\eta_\alpha$ is the $\alpha$th insightful ordinal; so the general existence question isn't interesting. However, I would be curious if there are any "natural" hierarchies of $M_\alpha$s which do satisfy this property; especially if their union is $H_{\omega_1}$ in some $V$ which is very far from $L$.</p>
[ { "answer_id": 259105, "score": 14, "text": "<p>It is a very nice question, but unfortunately, the answer is no. </p>\n\n<p><strong>Theorem.</strong> There are only countably many memorable ordinals.</p>\n\n<p><strong>Proof.</strong> Let $\\delta$ be a countable ordinal with $L_\\delta\\prec\nL_{\\omega_1}$. I claim that every memorable ordinal is less than\n$\\delta$.</p>\n\n<p>To see this, suppose that $\\alpha$ is memorable, as witnessed by\n$\\beta$. Fix an ordinal $\\gamma$ above both $\\delta$ and $\\beta$ with $L_\\gamma\\prec\nL_{\\omega_1}$. It follows that $L_\\delta\\prec L_\\gamma\\prec\nL_{\\omega_1}$. But since $\\beta\\leq\\gamma$, we know $\\alpha$ is\ndefinable in $L_\\gamma$ and hence (since there are no parameters)\ndefinable in $L_\\delta$. So $\\alpha&lt;\\delta$, as desired. <strong>QED</strong></p>\n" }, { "answer_id": 259500, "score": 3, "text": "<p>REPAIRED ARGUMENT</p>\n\n<p>The memorables are strictly bounded between the first $\\Sigma_2$-stable in $\\omega_1$ and the first $\\Sigma_3$-stable in $\\omega_1$. The least non-memorable is thus a $(\\Sigma_2\\wedge \\Pi_2)$-singleton.</p>\n\n<p>Let $\\delta_0$ be this first non-memorable. The previous erroneous argument yields a characterisation or perhaps a paraphrase of $\\delta_0$.</p>\n\n<p>For $\\beta&lt;\\omega_1$ let </p>\n\n<p>(A) $H(\\beta)$ be the Skolem Hull inside $L_\\beta$ of the empty set. </p>\n\n<p>(B) Let $\\omega_1(\\beta):= (\\omega_1)^{L_\\beta}$ if the latter is defined,\n $= \\beta$ otherwise.</p>\n\n<p>Then (i) $H(\\beta)$ is the set of pointwise definable objects in $L_\\beta$. (ii) $H(\\omega_1(\\beta))=$ $L_\\tau$ for some $\\tau\\leq \\omega_1(\\beta)$. (iii) For unboundedly many $\\beta$, $\\beta = \\omega_1(\\beta)$. </p>\n\n<p><em>Claim</em> Let $\\delta_1 =$ the least $\\delta&lt; \\omega_1$ so that for unboundedly many $\\beta$ $H(\\beta)=L_{\\delta_1}$. Then $\\delta_1=\\delta_0$.</p>\n\n<p>Proof: It is easy to argue that $\\delta_1$ is defined. Then, by definition $\\delta_1$ is not memorable (as $\\delta_1\n\\notin H(\\beta)$ for arbitrarily large $\\beta$). So it suffices to show that $\\tau&lt;\\delta_1 \\rightarrow \\tau$ is memorable.</p>\n\n<p>By definition, and countability, of $\\delta_1$:</p>\n\n<p>$\\exists \\beta_0\\forall \\beta&gt; \\beta_0\\forall\\tau&lt;\\delta_1\\,\\, H(\\beta)\\neq L_\\tau \\quad (*).$</p>\n\n<p>As $\\beta \\longrightarrow \\omega_1$ so does $\\omega_1(\\beta) \\longrightarrow \\omega_1$ non-decreasingly. Thus there is $\\beta_1&gt;\\beta_0$ so that:</p>\n\n<p>$\\forall\\beta&gt;\\beta_1\\,\\, \\omega_1\n(\\beta)&gt;\\beta_0.$</p>\n\n<p>Then, using (ii), for any $\\beta&gt;\\beta_1\\,\\, H(\\omega_1(\\beta))=L_\\gamma$ for some $\\gamma \\leq\\omega_1(\\beta)$ but by $(*)$ $\\delta_1\\leq\n\\gamma.$ Thus any $\\tau&lt; \\delta_1$ is pointwise definable in $H(\\omega_1(\\beta))$ and so (using a definition of $\\omega_1$), it is pointwise definable in $L_\\beta$. So all $\\tau$ less than $\\delta_1$ are memorable as required. $\\quad $ QED.</p>\n\n<p>So something similar would work if CH holds, or in CH models. Eg let $A\\subseteq\\omega_1$ be such that $H_{\\omega_1}=L_{\\omega_1}[A]$. Consider the least $\\delta$ with $L_\\delta[A\\cap \\delta]\\prec L_{\\omega_1}[A]$.</p>\n" }, { "answer_id": 259105, "score": 14, "text": "<p>It is a very nice question, but unfortunately, the answer is no. </p>\n\n<p><strong>Theorem.</strong> There are only countably many memorable ordinals.</p>\n\n<p><strong>Proof.</strong> Let $\\delta$ be a countable ordinal with $L_\\delta\\prec\nL_{\\omega_1}$. I claim that every memorable ordinal is less than\n$\\delta$.</p>\n\n<p>To see this, suppose that $\\alpha$ is memorable, as witnessed by\n$\\beta$. Fix an ordinal $\\gamma$ above both $\\delta$ and $\\beta$ with $L_\\gamma\\prec\nL_{\\omega_1}$. It follows that $L_\\delta\\prec L_\\gamma\\prec\nL_{\\omega_1}$. But since $\\beta\\leq\\gamma$, we know $\\alpha$ is\ndefinable in $L_\\gamma$ and hence (since there are no parameters)\ndefinable in $L_\\delta$. So $\\alpha&lt;\\delta$, as desired. <strong>QED</strong></p>\n" }, { "answer_id": 259500, "score": 3, "text": "<p>REPAIRED ARGUMENT</p>\n\n<p>The memorables are strictly bounded between the first $\\Sigma_2$-stable in $\\omega_1$ and the first $\\Sigma_3$-stable in $\\omega_1$. The least non-memorable is thus a $(\\Sigma_2\\wedge \\Pi_2)$-singleton.</p>\n\n<p>Let $\\delta_0$ be this first non-memorable. The previous erroneous argument yields a characterisation or perhaps a paraphrase of $\\delta_0$.</p>\n\n<p>For $\\beta&lt;\\omega_1$ let </p>\n\n<p>(A) $H(\\beta)$ be the Skolem Hull inside $L_\\beta$ of the empty set. </p>\n\n<p>(B) Let $\\omega_1(\\beta):= (\\omega_1)^{L_\\beta}$ if the latter is defined,\n $= \\beta$ otherwise.</p>\n\n<p>Then (i) $H(\\beta)$ is the set of pointwise definable objects in $L_\\beta$. (ii) $H(\\omega_1(\\beta))=$ $L_\\tau$ for some $\\tau\\leq \\omega_1(\\beta)$. (iii) For unboundedly many $\\beta$, $\\beta = \\omega_1(\\beta)$. </p>\n\n<p><em>Claim</em> Let $\\delta_1 =$ the least $\\delta&lt; \\omega_1$ so that for unboundedly many $\\beta$ $H(\\beta)=L_{\\delta_1}$. Then $\\delta_1=\\delta_0$.</p>\n\n<p>Proof: It is easy to argue that $\\delta_1$ is defined. Then, by definition $\\delta_1$ is not memorable (as $\\delta_1\n\\notin H(\\beta)$ for arbitrarily large $\\beta$). So it suffices to show that $\\tau&lt;\\delta_1 \\rightarrow \\tau$ is memorable.</p>\n\n<p>By definition, and countability, of $\\delta_1$:</p>\n\n<p>$\\exists \\beta_0\\forall \\beta&gt; \\beta_0\\forall\\tau&lt;\\delta_1\\,\\, H(\\beta)\\neq L_\\tau \\quad (*).$</p>\n\n<p>As $\\beta \\longrightarrow \\omega_1$ so does $\\omega_1(\\beta) \\longrightarrow \\omega_1$ non-decreasingly. Thus there is $\\beta_1&gt;\\beta_0$ so that:</p>\n\n<p>$\\forall\\beta&gt;\\beta_1\\,\\, \\omega_1\n(\\beta)&gt;\\beta_0.$</p>\n\n<p>Then, using (ii), for any $\\beta&gt;\\beta_1\\,\\, H(\\omega_1(\\beta))=L_\\gamma$ for some $\\gamma \\leq\\omega_1(\\beta)$ but by $(*)$ $\\delta_1\\leq\n\\gamma.$ Thus any $\\tau&lt; \\delta_1$ is pointwise definable in $H(\\omega_1(\\beta))$ and so (using a definition of $\\omega_1$), it is pointwise definable in $L_\\beta$. So all $\\tau$ less than $\\delta_1$ are memorable as required. $\\quad $ QED.</p>\n\n<p>So something similar would work if CH holds, or in CH models. Eg let $A\\subseteq\\omega_1$ be such that $H_{\\omega_1}=L_{\\omega_1}[A]$. Consider the least $\\delta$ with $L_\\delta[A\\cap \\delta]\\prec L_{\\omega_1}[A]$.</p>\n" }, { "answer_id": 259105, "score": 14, "text": "<p>It is a very nice question, but unfortunately, the answer is no. </p>\n\n<p><strong>Theorem.</strong> There are only countably many memorable ordinals.</p>\n\n<p><strong>Proof.</strong> Let $\\delta$ be a countable ordinal with $L_\\delta\\prec\nL_{\\omega_1}$. I claim that every memorable ordinal is less than\n$\\delta$.</p>\n\n<p>To see this, suppose that $\\alpha$ is memorable, as witnessed by\n$\\beta$. Fix an ordinal $\\gamma$ above both $\\delta$ and $\\beta$ with $L_\\gamma\\prec\nL_{\\omega_1}$. It follows that $L_\\delta\\prec L_\\gamma\\prec\nL_{\\omega_1}$. But since $\\beta\\leq\\gamma$, we know $\\alpha$ is\ndefinable in $L_\\gamma$ and hence (since there are no parameters)\ndefinable in $L_\\delta$. So $\\alpha&lt;\\delta$, as desired. <strong>QED</strong></p>\n" }, { "answer_id": 259500, "score": 3, "text": "<p>REPAIRED ARGUMENT</p>\n\n<p>The memorables are strictly bounded between the first $\\Sigma_2$-stable in $\\omega_1$ and the first $\\Sigma_3$-stable in $\\omega_1$. The least non-memorable is thus a $(\\Sigma_2\\wedge \\Pi_2)$-singleton.</p>\n\n<p>Let $\\delta_0$ be this first non-memorable. The previous erroneous argument yields a characterisation or perhaps a paraphrase of $\\delta_0$.</p>\n\n<p>For $\\beta&lt;\\omega_1$ let </p>\n\n<p>(A) $H(\\beta)$ be the Skolem Hull inside $L_\\beta$ of the empty set. </p>\n\n<p>(B) Let $\\omega_1(\\beta):= (\\omega_1)^{L_\\beta}$ if the latter is defined,\n $= \\beta$ otherwise.</p>\n\n<p>Then (i) $H(\\beta)$ is the set of pointwise definable objects in $L_\\beta$. (ii) $H(\\omega_1(\\beta))=$ $L_\\tau$ for some $\\tau\\leq \\omega_1(\\beta)$. (iii) For unboundedly many $\\beta$, $\\beta = \\omega_1(\\beta)$. </p>\n\n<p><em>Claim</em> Let $\\delta_1 =$ the least $\\delta&lt; \\omega_1$ so that for unboundedly many $\\beta$ $H(\\beta)=L_{\\delta_1}$. Then $\\delta_1=\\delta_0$.</p>\n\n<p>Proof: It is easy to argue that $\\delta_1$ is defined. Then, by definition $\\delta_1$ is not memorable (as $\\delta_1\n\\notin H(\\beta)$ for arbitrarily large $\\beta$). So it suffices to show that $\\tau&lt;\\delta_1 \\rightarrow \\tau$ is memorable.</p>\n\n<p>By definition, and countability, of $\\delta_1$:</p>\n\n<p>$\\exists \\beta_0\\forall \\beta&gt; \\beta_0\\forall\\tau&lt;\\delta_1\\,\\, H(\\beta)\\neq L_\\tau \\quad (*).$</p>\n\n<p>As $\\beta \\longrightarrow \\omega_1$ so does $\\omega_1(\\beta) \\longrightarrow \\omega_1$ non-decreasingly. Thus there is $\\beta_1&gt;\\beta_0$ so that:</p>\n\n<p>$\\forall\\beta&gt;\\beta_1\\,\\, \\omega_1\n(\\beta)&gt;\\beta_0.$</p>\n\n<p>Then, using (ii), for any $\\beta&gt;\\beta_1\\,\\, H(\\omega_1(\\beta))=L_\\gamma$ for some $\\gamma \\leq\\omega_1(\\beta)$ but by $(*)$ $\\delta_1\\leq\n\\gamma.$ Thus any $\\tau&lt; \\delta_1$ is pointwise definable in $H(\\omega_1(\\beta))$ and so (using a definition of $\\omega_1$), it is pointwise definable in $L_\\beta$. So all $\\tau$ less than $\\delta_1$ are memorable as required. $\\quad $ QED.</p>\n\n<p>So something similar would work if CH holds, or in CH models. Eg let $A\\subseteq\\omega_1$ be such that $H_{\\omega_1}=L_{\\omega_1}[A]$. Consider the least $\\delta$ with $L_\\delta[A\\cap \\delta]\\prec L_{\\omega_1}[A]$.</p>\n" } ]
Question: Memorable ordinals Answer 1: <p>It is a very nice question, but unfortunately, the answer is no. </p> <p><strong>Theorem.</strong> There are only countably many memorable ordinals.</p> <p><strong>Proof.</strong> Let $\delta$ be a countable ordinal with $L_\delta\prec L_{\omega_1}$. I claim that every memorable ordinal is less than $\delta$.</p> <p>To see this, suppose that $\alpha$ is memorable, as witnessed by $\beta$. Fix an ordinal $\gamma$ above both $\delta$ and $\beta$ with $L_\gamma\prec L_{\omega_1}$. It follows that $L_\delta\prec L_\gamma\prec L_{\omega_1}$. But since $\beta\leq\gamma$, we know $\alpha$ is definable in $L_\gamma$ and hence (since there are no parameters) definable in $L_\delta$. So $\alpha&lt;\delta$, as desired. <strong>QED</strong></p> Answer 2: <p>REPAIRED ARGUMENT</p> <p>The memorables are strictly bounded between the first $\Sigma_2$-stable in $\omega_1$ and the first $\Sigma_3$-stable in $\omega_1$. The least non-memorable is thus a $(\Sigma_2\wedge \Pi_2)$-singleton.</p> <p>Let $\delta_0$ be this first non-memorable. The previous erroneous argument yields a characterisation or perhaps a paraphrase of $\delta_0$.</p> <p>For $\beta&lt;\omega_1$ let </p> <p>(A) $H(\beta)$ be the Skolem Hull inside $L_\beta$ of the empty set. </p> <p>(B) Let $\omega_1(\beta):= (\omega_1)^{L_\beta}$ if the latter is defined, $= \beta$ otherwise.</p> <p>Then (i) $H(\beta)$ is the set of pointwise definable objects in $L_\beta$. (ii) $H(\omega_1(\beta))=$ $L_\tau$ for some $\tau\leq \omega_1(\beta)$. (iii) For unboundedly many $\beta$, $\beta = \omega_1(\beta)$. </p> <p><em>Claim</em> Let $\delta_1 =$ the least $\delta&lt; \omega_1$ so that for unboundedly many $\beta$ $H(\beta)=L_{\delta_1}$. Then $\delta_1=\delta_0$.</p> <p>Proof: It is easy to argue that $\delta_1$ is defined. Then, by definition $\delta_1$ is not memorable (as $\delta_1 \notin H(\beta)$ for arbitrarily large $\beta$). So it suffices to show that $\tau&lt;\delta_1 \rightarrow \tau$ is memorable.</p> <p>By definition, and countability, of $\delta_1$:</p> <p>$\exists \beta_0\forall \beta&gt; \beta_0\forall\tau&lt;\delta_1\,\, H(\beta)\neq L_\tau \quad (*).$</p> <p>As $\beta \longrightarrow \omega_1$ so does $\omega_1(\beta) \longrightarrow \omega_1$ non-decreasingly. Thus there is $\beta_1&gt;\beta_0$ so that:</p> <p>$\forall\beta&gt;\beta_1\,\, \omega_1 (\beta)&gt;\beta_0.$</p> <p>Then, using (ii), for any $\beta&gt;\beta_1\,\, H(\omega_1(\beta))=L_\gamma$ for some $\gamma \leq\omega_1(\beta)$ but by $(*)$ $\delta_1\leq \gamma.$ Thus any $\tau&lt; \delta_1$ is pointwise definable in $H(\omega_1(\beta))$ and so (using a definition of $\omega_1$), it is pointwise definable in $L_\beta$. So all $\tau$ less than $\delta_1$ are memorable as required. $\quad $ QED.</p> <p>So something similar would work if CH holds, or in CH models. Eg let $A\subseteq\omega_1$ be such that $H_{\omega_1}=L_{\omega_1}[A]$. Consider the least $\delta$ with $L_\delta[A\cap \delta]\prec L_{\omega_1}[A]$.</p> Answer 3: <p>It is a very nice question, but unfortunately, the answer is no. </p> <p><strong>Theorem.</strong> There are only countably many memorable ordinals.</p> <p><strong>Proof.</strong> Let $\delta$ be a countable ordinal with $L_\delta\prec L_{\omega_1}$. I claim that every memorable ordinal is less than $\delta$.</p> <p>To see this, suppose that $\alpha$ is memorable, as witnessed by $\beta$. Fix an ordinal $\gamma$ above both $\delta$ and $\beta$ with $L_\gamma\prec L_{\omega_1}$. It follows that $L_\delta\prec L_\gamma\prec L_{\omega_1}$. But since $\beta\leq\gamma$, we know $\alpha$ is definable in $L_\gamma$ and hence (since there are no parameters) definable in $L_\delta$. So $\alpha&lt;\delta$, as desired. <strong>QED</strong></p> Answer 4: <p>REPAIRED ARGUMENT</p> <p>The memorables are strictly bounded between the first $\Sigma_2$-stable in $\omega_1$ and the first $\Sigma_3$-stable in $\omega_1$. The least non-memorable is thus a $(\Sigma_2\wedge \Pi_2)$-singleton.</p> <p>Let $\delta_0$ be this first non-memorable. The previous erroneous argument yields a characterisation or perhaps a paraphrase of $\delta_0$.</p> <p>For $\beta&lt;\omega_1$ let </p> <p>(A) $H(\beta)$ be the Skolem Hull inside $L_\beta$ of the empty set. </p> <p>(B) Let $\omega_1(\beta):= (\omega_1)^{L_\beta}$ if the latter is defined, $= \beta$ otherwise.</p> <p>Then (i) $H(\beta)$ is the set of pointwise definable objects in $L_\beta$. (ii) $H(\omega_1(\beta))=$ $L_\tau$ for some $\tau\leq \omega_1(\beta)$. (iii) For unboundedly many $\beta$, $\beta = \omega_1(\beta)$. </p> <p><em>Claim</em> Let $\delta_1 =$ the least $\delta&lt; \omega_1$ so that for unboundedly many $\beta$ $H(\beta)=L_{\delta_1}$. Then $\delta_1=\delta_0$.</p> <p>Proof: It is easy to argue that $\delta_1$ is defined. Then, by definition $\delta_1$ is not memorable (as $\delta_1 \notin H(\beta)$ for arbitrarily large $\beta$). So it suffices to show that $\tau&lt;\delta_1 \rightarrow \tau$ is memorable.</p> <p>By definition, and countability, of $\delta_1$:</p> <p>$\exists \beta_0\forall \beta&gt; \beta_0\forall\tau&lt;\delta_1\,\, H(\beta)\neq L_\tau \quad (*).$</p> <p>As $\beta \longrightarrow \omega_1$ so does $\omega_1(\beta) \longrightarrow \omega_1$ non-decreasingly. Thus there is $\beta_1&gt;\beta_0$ so that:</p> <p>$\forall\beta&gt;\beta_1\,\, \omega_1 (\beta)&gt;\beta_0.$</p> <p>Then, using (ii), for any $\beta&gt;\beta_1\,\, H(\omega_1(\beta))=L_\gamma$ for some $\gamma \leq\omega_1(\beta)$ but by $(*)$ $\delta_1\leq \gamma.$ Thus any $\tau&lt; \delta_1$ is pointwise definable in $H(\omega_1(\beta))$ and so (using a definition of $\omega_1$), it is pointwise definable in $L_\beta$. So all $\tau$ less than $\delta_1$ are memorable as required. $\quad $ QED.</p> <p>So something similar would work if CH holds, or in CH models. Eg let $A\subseteq\omega_1$ be such that $H_{\omega_1}=L_{\omega_1}[A]$. Consider the least $\delta$ with $L_\delta[A\cap \delta]\prec L_{\omega_1}[A]$.</p> Answer 5: <p>It is a very nice question, but unfortunately, the answer is no. </p> <p><strong>Theorem.</strong> There are only countably many memorable ordinals.</p> <p><strong>Proof.</strong> Let $\delta$ be a countable ordinal with $L_\delta\prec L_{\omega_1}$. I claim that every memorable ordinal is less than $\delta$.</p> <p>To see this, suppose that $\alpha$ is memorable, as witnessed by $\beta$. Fix an ordinal $\gamma$ above both $\delta$ and $\beta$ with $L_\gamma\prec L_{\omega_1}$. It follows that $L_\delta\prec L_\gamma\prec L_{\omega_1}$. But since $\beta\leq\gamma$, we know $\alpha$ is definable in $L_\gamma$ and hence (since there are no parameters) definable in $L_\delta$. So $\alpha&lt;\delta$, as desired. <strong>QED</strong></p> Answer 6: <p>REPAIRED ARGUMENT</p> <p>The memorables are strictly bounded between the first $\Sigma_2$-stable in $\omega_1$ and the first $\Sigma_3$-stable in $\omega_1$. The least non-memorable is thus a $(\Sigma_2\wedge \Pi_2)$-singleton.</p> <p>Let $\delta_0$ be this first non-memorable. The previous erroneous argument yields a characterisation or perhaps a paraphrase of $\delta_0$.</p> <p>For $\beta&lt;\omega_1$ let </p> <p>(A) $H(\beta)$ be the Skolem Hull inside $L_\beta$ of the empty set. </p> <p>(B) Let $\omega_1(\beta):= (\omega_1)^{L_\beta}$ if the latter is defined, $= \beta$ otherwise.</p> <p>Then (i) $H(\beta)$ is the set of pointwise definable objects in $L_\beta$. (ii) $H(\omega_1(\beta))=$ $L_\tau$ for some $\tau\leq \omega_1(\beta)$. (iii) For unboundedly many $\beta$, $\beta = \omega_1(\beta)$. </p> <p><em>Claim</em> Let $\delta_1 =$ the least $\delta&lt; \omega_1$ so that for unboundedly many $\beta$ $H(\beta)=L_{\delta_1}$. Then $\delta_1=\delta_0$.</p> <p>Proof: It is easy to argue that $\delta_1$ is defined. Then, by definition $\delta_1$ is not memorable (as $\delta_1 \notin H(\beta)$ for arbitrarily large $\beta$). So it suffices to show that $\tau&lt;\delta_1 \rightarrow \tau$ is memorable.</p> <p>By definition, and countability, of $\delta_1$:</p> <p>$\exists \beta_0\forall \beta&gt; \beta_0\forall\tau&lt;\delta_1\,\, H(\beta)\neq L_\tau \quad (*).$</p> <p>As $\beta \longrightarrow \omega_1$ so does $\omega_1(\beta) \longrightarrow \omega_1$ non-decreasingly. Thus there is $\beta_1&gt;\beta_0$ so that:</p> <p>$\forall\beta&gt;\beta_1\,\, \omega_1 (\beta)&gt;\beta_0.$</p> <p>Then, using (ii), for any $\beta&gt;\beta_1\,\, H(\omega_1(\beta))=L_\gamma$ for some $\gamma \leq\omega_1(\beta)$ but by $(*)$ $\delta_1\leq \gamma.$ Thus any $\tau&lt; \delta_1$ is pointwise definable in $H(\omega_1(\beta))$ and so (using a definition of $\omega_1$), it is pointwise definable in $L_\beta$. So all $\tau$ less than $\delta_1$ are memorable as required. $\quad $ QED.</p> <p>So something similar would work if CH holds, or in CH models. Eg let $A\subseteq\omega_1$ be such that $H_{\omega_1}=L_{\omega_1}[A]$. Consider the least $\delta$ with $L_\delta[A\cap \delta]\prec L_{\omega_1}[A]$.</p>
Computation of $\pi_4$ of simple Lie groups
259,487
13
https://mathoverflow.net/questions/259487/computation-of-pi-4-of-simple-lie-groups
<p>Below we assume any simple Lie group $G$ to be simply connected.</p> <p>$\pi_3(G)=\mathbb{Z}$ for any simple Lie group $G$ and there is a uniform proof for that.</p> <p>Now the textbooks say $\pi_4(G)$ is trivial except for $G=Sp(n)$, for which it is $\mathbb{Z}/2\mathbb{Z}$. </p> <p>My question is the following: is there a uniform way to derive this fact, in such a way to show which property of the symplectic group is so special?</p>
[ { "answer_id": 259527, "score": 13, "text": "<p>As a partial answer, here's at least a uniform statement, which can be found as Theorem 3.10 in Mimura's survey \"Homotopy theory of Lie groups\" in the Handbook of Algebraic Topology: </p>\n\n<blockquote>\n <p>Let $G$ be a compact, connected, simply connected, simple Lie group, $T$ a maximal torus of $G$, $\\mathbb{R}^r$ its universal covering and $\\Gamma$ the inverse image of the identity in $\\mathbb{R}^r$. Let $a$ be the dominant root with respect to some lexicographic order of the roots of $G$. Then $\\pi_4(G)$ is isomorphic to $\\mathbb{Z}/2\\mathbb{Z}$ if the hyperplane $a=1$ contains a point of $\\Gamma$ and $0$ otherwise. </p>\n</blockquote>\n\n<p>(In the survey article, the statement has a $\\mathbb{Z}$ where I wrote $\\mathbb{Z}/2\\mathbb{Z}$ but that would appear to be a typo.)</p>\n\n<p>Mimura refers to the paper</p>\n\n<ul>\n<li>R. Bott and H. Samelson. Applications of the theory of Morse to symmetric spaces. Amer. J. Math. 80 (1958), 964-1029. </li>\n</ul>\n\n<p>I can't check right now, but the uniform statement suggests a uniform proof, relating the fourth homotopy group to information on the root system. It would seem that the difference that singles out the symplectic groups is that the long roots are divisible (by 2 incidentally) in the character group.</p>\n" }, { "answer_id": 259524, "score": 12, "text": "<p>Since you are asking a question about $\\pi_4$, lets use a little homotopy theory to think about this, and for starters, we can just ponder compact Lie groups. </p>\n\n<p>Consider first $G=Sp(1)=SU(2)=S^3$. It is well known that $\\pi_4(S^3) = \\mathbb Z/2$. </p>\n\n<p>Now lets compare and contrast $SU(3)$ with $Sp(2)$. $SU(3)/SU(2) = S^5$, while $Sp(2)/Sp(1) = S^7$, and these lead to fibration sequences\n$ S^3 \\rightarrow SU(3) \\rightarrow S^5$\nand \n$ S^3 \\rightarrow Sp(2) \\rightarrow S^7$. The associated long exact sequences of homotopy groups make it obvious that $\\pi_4(Sp(2)) = \\mathbb Z/2$; more surprising from this perspective is that $\\pi_4(SU(3)) = 0$.</p>\n\n<p>In general, well chosen (= highly connected) homogeneous spaces $G/H$ let one get at $\\pi_*(G)$ from knowledge of $\\pi_*(H)$ for $*$ not too large. For example, $G_2/SU(3) = S^6$, so $\\pi_4(G_2) = 0$.</p>\n\n<p>Finally, here is a general method: Let $G\\langle 3\\rangle$ be the fiber of the fundamental class $G \\rightarrow K(\\mathbb Z,3)$. Then $\\pi_4(G) = \\pi_4(G\\langle 3\\rangle) = H_4(G\\langle 3\\rangle;\\mathbb Z)$. Then one can try to get at this homology group using the Serre spectral sequence associated to the fibration $K(\\mathbb Z,2) \\rightarrow G\\langle 3 \\rangle \\rightarrow G$. </p>\n\n<p>In various papers, M. Mimura has done much analysis of this sort, and in particular computes the first couple dozen homotopy groups of all of the exceptional Lie groups.</p>\n" }, { "answer_id": 259600, "score": 3, "text": "<p>There is actually a collection of more general statements that detail the structure of the $\\pi_4$.</p>\n\n<p>Firstly it was Browder [The Cohomology of Covering Spaces of H-Spaces] who proved that a simply connected finite H-space is 2-connected. Clark [On $\\pi_3$ of Finite Dimensional H-Spaces] showed that the first non-vanishing homotopy group of a finite loop space appeared in degrees 1 or 3 (the result that $\\pi_2=0$ for a simply connected finite H-space also appears in Browders [Torsion in H-Spaces], see also Lin's [The First Homotopy Group of a Finite H-Space]). Kane and Hubbuck [On $\\pi_3$ of a Finite H-Space] prove that $\\pi_3$ is torsion free, building on previous work of Thomas and Lin which required stricter assumptions (the Lie group $E_8$ does not fall under Thomas's requirements, for instance).</p>\n\n<p>Finally Hubbuck [Homotopy Groups of Finite H-Spaces] studied $\\pi_4$, showing that it was a direct sum of finite groups of order 2. The proof uses JHC Whitehead's Certain Exact Sequence (see Baues [Homotopy Type and Homology] for a good discussion), identifying the groups that appear within using the work of others. The end result is an exact sequence</p>\n\n<p>$H_5(X;\\mathbb{Z})\\xrightarrow{\\rho_2} H_5(X;\\mathbb{Z}_2)\\xrightarrow{Sq^2_*}H_3(X;\\mathbb{Z}_2)\\rightarrow \\pi_4X\\rightarrow0$</p>\n\n<p>And we know that $H_3(X;\\mathbb{Z}_2)=\\oplus\\,\\mathbb{Z}_2$ from the Hurewicz theorem and the work of Kane and Hubbuck so the result on $\\pi_4X$ follows.</p>\n\n<p>Harper claims that if $X$ is simply connected and $H_*(\\Omega X;\\mathbb{Z})$ is torsion free then the result follows from the previous work of Bott and Samelson [Applications of the Theory of Morse to Symmetric Spaces].</p>\n" }, { "answer_id": 259527, "score": 13, "text": "<p>As a partial answer, here's at least a uniform statement, which can be found as Theorem 3.10 in Mimura's survey \"Homotopy theory of Lie groups\" in the Handbook of Algebraic Topology: </p>\n\n<blockquote>\n <p>Let $G$ be a compact, connected, simply connected, simple Lie group, $T$ a maximal torus of $G$, $\\mathbb{R}^r$ its universal covering and $\\Gamma$ the inverse image of the identity in $\\mathbb{R}^r$. Let $a$ be the dominant root with respect to some lexicographic order of the roots of $G$. Then $\\pi_4(G)$ is isomorphic to $\\mathbb{Z}/2\\mathbb{Z}$ if the hyperplane $a=1$ contains a point of $\\Gamma$ and $0$ otherwise. </p>\n</blockquote>\n\n<p>(In the survey article, the statement has a $\\mathbb{Z}$ where I wrote $\\mathbb{Z}/2\\mathbb{Z}$ but that would appear to be a typo.)</p>\n\n<p>Mimura refers to the paper</p>\n\n<ul>\n<li>R. Bott and H. Samelson. Applications of the theory of Morse to symmetric spaces. Amer. J. Math. 80 (1958), 964-1029. </li>\n</ul>\n\n<p>I can't check right now, but the uniform statement suggests a uniform proof, relating the fourth homotopy group to information on the root system. It would seem that the difference that singles out the symplectic groups is that the long roots are divisible (by 2 incidentally) in the character group.</p>\n" }, { "answer_id": 259524, "score": 12, "text": "<p>Since you are asking a question about $\\pi_4$, lets use a little homotopy theory to think about this, and for starters, we can just ponder compact Lie groups. </p>\n\n<p>Consider first $G=Sp(1)=SU(2)=S^3$. It is well known that $\\pi_4(S^3) = \\mathbb Z/2$. </p>\n\n<p>Now lets compare and contrast $SU(3)$ with $Sp(2)$. $SU(3)/SU(2) = S^5$, while $Sp(2)/Sp(1) = S^7$, and these lead to fibration sequences\n$ S^3 \\rightarrow SU(3) \\rightarrow S^5$\nand \n$ S^3 \\rightarrow Sp(2) \\rightarrow S^7$. The associated long exact sequences of homotopy groups make it obvious that $\\pi_4(Sp(2)) = \\mathbb Z/2$; more surprising from this perspective is that $\\pi_4(SU(3)) = 0$.</p>\n\n<p>In general, well chosen (= highly connected) homogeneous spaces $G/H$ let one get at $\\pi_*(G)$ from knowledge of $\\pi_*(H)$ for $*$ not too large. For example, $G_2/SU(3) = S^6$, so $\\pi_4(G_2) = 0$.</p>\n\n<p>Finally, here is a general method: Let $G\\langle 3\\rangle$ be the fiber of the fundamental class $G \\rightarrow K(\\mathbb Z,3)$. Then $\\pi_4(G) = \\pi_4(G\\langle 3\\rangle) = H_4(G\\langle 3\\rangle;\\mathbb Z)$. Then one can try to get at this homology group using the Serre spectral sequence associated to the fibration $K(\\mathbb Z,2) \\rightarrow G\\langle 3 \\rangle \\rightarrow G$. </p>\n\n<p>In various papers, M. Mimura has done much analysis of this sort, and in particular computes the first couple dozen homotopy groups of all of the exceptional Lie groups.</p>\n" }, { "answer_id": 259600, "score": 3, "text": "<p>There is actually a collection of more general statements that detail the structure of the $\\pi_4$.</p>\n\n<p>Firstly it was Browder [The Cohomology of Covering Spaces of H-Spaces] who proved that a simply connected finite H-space is 2-connected. Clark [On $\\pi_3$ of Finite Dimensional H-Spaces] showed that the first non-vanishing homotopy group of a finite loop space appeared in degrees 1 or 3 (the result that $\\pi_2=0$ for a simply connected finite H-space also appears in Browders [Torsion in H-Spaces], see also Lin's [The First Homotopy Group of a Finite H-Space]). Kane and Hubbuck [On $\\pi_3$ of a Finite H-Space] prove that $\\pi_3$ is torsion free, building on previous work of Thomas and Lin which required stricter assumptions (the Lie group $E_8$ does not fall under Thomas's requirements, for instance).</p>\n\n<p>Finally Hubbuck [Homotopy Groups of Finite H-Spaces] studied $\\pi_4$, showing that it was a direct sum of finite groups of order 2. The proof uses JHC Whitehead's Certain Exact Sequence (see Baues [Homotopy Type and Homology] for a good discussion), identifying the groups that appear within using the work of others. The end result is an exact sequence</p>\n\n<p>$H_5(X;\\mathbb{Z})\\xrightarrow{\\rho_2} H_5(X;\\mathbb{Z}_2)\\xrightarrow{Sq^2_*}H_3(X;\\mathbb{Z}_2)\\rightarrow \\pi_4X\\rightarrow0$</p>\n\n<p>And we know that $H_3(X;\\mathbb{Z}_2)=\\oplus\\,\\mathbb{Z}_2$ from the Hurewicz theorem and the work of Kane and Hubbuck so the result on $\\pi_4X$ follows.</p>\n\n<p>Harper claims that if $X$ is simply connected and $H_*(\\Omega X;\\mathbb{Z})$ is torsion free then the result follows from the previous work of Bott and Samelson [Applications of the Theory of Morse to Symmetric Spaces].</p>\n" }, { "answer_id": 259527, "score": 13, "text": "<p>As a partial answer, here's at least a uniform statement, which can be found as Theorem 3.10 in Mimura's survey \"Homotopy theory of Lie groups\" in the Handbook of Algebraic Topology: </p>\n\n<blockquote>\n <p>Let $G$ be a compact, connected, simply connected, simple Lie group, $T$ a maximal torus of $G$, $\\mathbb{R}^r$ its universal covering and $\\Gamma$ the inverse image of the identity in $\\mathbb{R}^r$. Let $a$ be the dominant root with respect to some lexicographic order of the roots of $G$. Then $\\pi_4(G)$ is isomorphic to $\\mathbb{Z}/2\\mathbb{Z}$ if the hyperplane $a=1$ contains a point of $\\Gamma$ and $0$ otherwise. </p>\n</blockquote>\n\n<p>(In the survey article, the statement has a $\\mathbb{Z}$ where I wrote $\\mathbb{Z}/2\\mathbb{Z}$ but that would appear to be a typo.)</p>\n\n<p>Mimura refers to the paper</p>\n\n<ul>\n<li>R. Bott and H. Samelson. Applications of the theory of Morse to symmetric spaces. Amer. J. Math. 80 (1958), 964-1029. </li>\n</ul>\n\n<p>I can't check right now, but the uniform statement suggests a uniform proof, relating the fourth homotopy group to information on the root system. It would seem that the difference that singles out the symplectic groups is that the long roots are divisible (by 2 incidentally) in the character group.</p>\n" }, { "answer_id": 259524, "score": 12, "text": "<p>Since you are asking a question about $\\pi_4$, lets use a little homotopy theory to think about this, and for starters, we can just ponder compact Lie groups. </p>\n\n<p>Consider first $G=Sp(1)=SU(2)=S^3$. It is well known that $\\pi_4(S^3) = \\mathbb Z/2$. </p>\n\n<p>Now lets compare and contrast $SU(3)$ with $Sp(2)$. $SU(3)/SU(2) = S^5$, while $Sp(2)/Sp(1) = S^7$, and these lead to fibration sequences\n$ S^3 \\rightarrow SU(3) \\rightarrow S^5$\nand \n$ S^3 \\rightarrow Sp(2) \\rightarrow S^7$. The associated long exact sequences of homotopy groups make it obvious that $\\pi_4(Sp(2)) = \\mathbb Z/2$; more surprising from this perspective is that $\\pi_4(SU(3)) = 0$.</p>\n\n<p>In general, well chosen (= highly connected) homogeneous spaces $G/H$ let one get at $\\pi_*(G)$ from knowledge of $\\pi_*(H)$ for $*$ not too large. For example, $G_2/SU(3) = S^6$, so $\\pi_4(G_2) = 0$.</p>\n\n<p>Finally, here is a general method: Let $G\\langle 3\\rangle$ be the fiber of the fundamental class $G \\rightarrow K(\\mathbb Z,3)$. Then $\\pi_4(G) = \\pi_4(G\\langle 3\\rangle) = H_4(G\\langle 3\\rangle;\\mathbb Z)$. Then one can try to get at this homology group using the Serre spectral sequence associated to the fibration $K(\\mathbb Z,2) \\rightarrow G\\langle 3 \\rangle \\rightarrow G$. </p>\n\n<p>In various papers, M. Mimura has done much analysis of this sort, and in particular computes the first couple dozen homotopy groups of all of the exceptional Lie groups.</p>\n" }, { "answer_id": 259600, "score": 3, "text": "<p>There is actually a collection of more general statements that detail the structure of the $\\pi_4$.</p>\n\n<p>Firstly it was Browder [The Cohomology of Covering Spaces of H-Spaces] who proved that a simply connected finite H-space is 2-connected. Clark [On $\\pi_3$ of Finite Dimensional H-Spaces] showed that the first non-vanishing homotopy group of a finite loop space appeared in degrees 1 or 3 (the result that $\\pi_2=0$ for a simply connected finite H-space also appears in Browders [Torsion in H-Spaces], see also Lin's [The First Homotopy Group of a Finite H-Space]). Kane and Hubbuck [On $\\pi_3$ of a Finite H-Space] prove that $\\pi_3$ is torsion free, building on previous work of Thomas and Lin which required stricter assumptions (the Lie group $E_8$ does not fall under Thomas's requirements, for instance).</p>\n\n<p>Finally Hubbuck [Homotopy Groups of Finite H-Spaces] studied $\\pi_4$, showing that it was a direct sum of finite groups of order 2. The proof uses JHC Whitehead's Certain Exact Sequence (see Baues [Homotopy Type and Homology] for a good discussion), identifying the groups that appear within using the work of others. The end result is an exact sequence</p>\n\n<p>$H_5(X;\\mathbb{Z})\\xrightarrow{\\rho_2} H_5(X;\\mathbb{Z}_2)\\xrightarrow{Sq^2_*}H_3(X;\\mathbb{Z}_2)\\rightarrow \\pi_4X\\rightarrow0$</p>\n\n<p>And we know that $H_3(X;\\mathbb{Z}_2)=\\oplus\\,\\mathbb{Z}_2$ from the Hurewicz theorem and the work of Kane and Hubbuck so the result on $\\pi_4X$ follows.</p>\n\n<p>Harper claims that if $X$ is simply connected and $H_*(\\Omega X;\\mathbb{Z})$ is torsion free then the result follows from the previous work of Bott and Samelson [Applications of the Theory of Morse to Symmetric Spaces].</p>\n" } ]
Question: Computation of $\pi_4$ of simple Lie groups Answer 1: <p>As a partial answer, here's at least a uniform statement, which can be found as Theorem 3.10 in Mimura's survey "Homotopy theory of Lie groups" in the Handbook of Algebraic Topology: </p> <blockquote> <p>Let $G$ be a compact, connected, simply connected, simple Lie group, $T$ a maximal torus of $G$, $\mathbb{R}^r$ its universal covering and $\Gamma$ the inverse image of the identity in $\mathbb{R}^r$. Let $a$ be the dominant root with respect to some lexicographic order of the roots of $G$. Then $\pi_4(G)$ is isomorphic to $\mathbb{Z}/2\mathbb{Z}$ if the hyperplane $a=1$ contains a point of $\Gamma$ and $0$ otherwise. </p> </blockquote> <p>(In the survey article, the statement has a $\mathbb{Z}$ where I wrote $\mathbb{Z}/2\mathbb{Z}$ but that would appear to be a typo.)</p> <p>Mimura refers to the paper</p> <ul> <li>R. Bott and H. Samelson. Applications of the theory of Morse to symmetric spaces. Amer. J. Math. 80 (1958), 964-1029. </li> </ul> <p>I can't check right now, but the uniform statement suggests a uniform proof, relating the fourth homotopy group to information on the root system. It would seem that the difference that singles out the symplectic groups is that the long roots are divisible (by 2 incidentally) in the character group.</p> Answer 2: <p>Since you are asking a question about $\pi_4$, lets use a little homotopy theory to think about this, and for starters, we can just ponder compact Lie groups. </p> <p>Consider first $G=Sp(1)=SU(2)=S^3$. It is well known that $\pi_4(S^3) = \mathbb Z/2$. </p> <p>Now lets compare and contrast $SU(3)$ with $Sp(2)$. $SU(3)/SU(2) = S^5$, while $Sp(2)/Sp(1) = S^7$, and these lead to fibration sequences $ S^3 \rightarrow SU(3) \rightarrow S^5$ and $ S^3 \rightarrow Sp(2) \rightarrow S^7$. The associated long exact sequences of homotopy groups make it obvious that $\pi_4(Sp(2)) = \mathbb Z/2$; more surprising from this perspective is that $\pi_4(SU(3)) = 0$.</p> <p>In general, well chosen (= highly connected) homogeneous spaces $G/H$ let one get at $\pi_*(G)$ from knowledge of $\pi_*(H)$ for $*$ not too large. For example, $G_2/SU(3) = S^6$, so $\pi_4(G_2) = 0$.</p> <p>Finally, here is a general method: Let $G\langle 3\rangle$ be the fiber of the fundamental class $G \rightarrow K(\mathbb Z,3)$. Then $\pi_4(G) = \pi_4(G\langle 3\rangle) = H_4(G\langle 3\rangle;\mathbb Z)$. Then one can try to get at this homology group using the Serre spectral sequence associated to the fibration $K(\mathbb Z,2) \rightarrow G\langle 3 \rangle \rightarrow G$. </p> <p>In various papers, M. Mimura has done much analysis of this sort, and in particular computes the first couple dozen homotopy groups of all of the exceptional Lie groups.</p> Answer 3: <p>There is actually a collection of more general statements that detail the structure of the $\pi_4$.</p> <p>Firstly it was Browder [The Cohomology of Covering Spaces of H-Spaces] who proved that a simply connected finite H-space is 2-connected. Clark [On $\pi_3$ of Finite Dimensional H-Spaces] showed that the first non-vanishing homotopy group of a finite loop space appeared in degrees 1 or 3 (the result that $\pi_2=0$ for a simply connected finite H-space also appears in Browders [Torsion in H-Spaces], see also Lin's [The First Homotopy Group of a Finite H-Space]). Kane and Hubbuck [On $\pi_3$ of a Finite H-Space] prove that $\pi_3$ is torsion free, building on previous work of Thomas and Lin which required stricter assumptions (the Lie group $E_8$ does not fall under Thomas's requirements, for instance).</p> <p>Finally Hubbuck [Homotopy Groups of Finite H-Spaces] studied $\pi_4$, showing that it was a direct sum of finite groups of order 2. The proof uses JHC Whitehead's Certain Exact Sequence (see Baues [Homotopy Type and Homology] for a good discussion), identifying the groups that appear within using the work of others. The end result is an exact sequence</p> <p>$H_5(X;\mathbb{Z})\xrightarrow{\rho_2} H_5(X;\mathbb{Z}_2)\xrightarrow{Sq^2_*}H_3(X;\mathbb{Z}_2)\rightarrow \pi_4X\rightarrow0$</p> <p>And we know that $H_3(X;\mathbb{Z}_2)=\oplus\,\mathbb{Z}_2$ from the Hurewicz theorem and the work of Kane and Hubbuck so the result on $\pi_4X$ follows.</p> <p>Harper claims that if $X$ is simply connected and $H_*(\Omega X;\mathbb{Z})$ is torsion free then the result follows from the previous work of Bott and Samelson [Applications of the Theory of Morse to Symmetric Spaces].</p> Answer 4: <p>As a partial answer, here's at least a uniform statement, which can be found as Theorem 3.10 in Mimura's survey "Homotopy theory of Lie groups" in the Handbook of Algebraic Topology: </p> <blockquote> <p>Let $G$ be a compact, connected, simply connected, simple Lie group, $T$ a maximal torus of $G$, $\mathbb{R}^r$ its universal covering and $\Gamma$ the inverse image of the identity in $\mathbb{R}^r$. Let $a$ be the dominant root with respect to some lexicographic order of the roots of $G$. Then $\pi_4(G)$ is isomorphic to $\mathbb{Z}/2\mathbb{Z}$ if the hyperplane $a=1$ contains a point of $\Gamma$ and $0$ otherwise. </p> </blockquote> <p>(In the survey article, the statement has a $\mathbb{Z}$ where I wrote $\mathbb{Z}/2\mathbb{Z}$ but that would appear to be a typo.)</p> <p>Mimura refers to the paper</p> <ul> <li>R. Bott and H. Samelson. Applications of the theory of Morse to symmetric spaces. Amer. J. Math. 80 (1958), 964-1029. </li> </ul> <p>I can't check right now, but the uniform statement suggests a uniform proof, relating the fourth homotopy group to information on the root system. It would seem that the difference that singles out the symplectic groups is that the long roots are divisible (by 2 incidentally) in the character group.</p> Answer 5: <p>Since you are asking a question about $\pi_4$, lets use a little homotopy theory to think about this, and for starters, we can just ponder compact Lie groups. </p> <p>Consider first $G=Sp(1)=SU(2)=S^3$. It is well known that $\pi_4(S^3) = \mathbb Z/2$. </p> <p>Now lets compare and contrast $SU(3)$ with $Sp(2)$. $SU(3)/SU(2) = S^5$, while $Sp(2)/Sp(1) = S^7$, and these lead to fibration sequences $ S^3 \rightarrow SU(3) \rightarrow S^5$ and $ S^3 \rightarrow Sp(2) \rightarrow S^7$. The associated long exact sequences of homotopy groups make it obvious that $\pi_4(Sp(2)) = \mathbb Z/2$; more surprising from this perspective is that $\pi_4(SU(3)) = 0$.</p> <p>In general, well chosen (= highly connected) homogeneous spaces $G/H$ let one get at $\pi_*(G)$ from knowledge of $\pi_*(H)$ for $*$ not too large. For example, $G_2/SU(3) = S^6$, so $\pi_4(G_2) = 0$.</p> <p>Finally, here is a general method: Let $G\langle 3\rangle$ be the fiber of the fundamental class $G \rightarrow K(\mathbb Z,3)$. Then $\pi_4(G) = \pi_4(G\langle 3\rangle) = H_4(G\langle 3\rangle;\mathbb Z)$. Then one can try to get at this homology group using the Serre spectral sequence associated to the fibration $K(\mathbb Z,2) \rightarrow G\langle 3 \rangle \rightarrow G$. </p> <p>In various papers, M. Mimura has done much analysis of this sort, and in particular computes the first couple dozen homotopy groups of all of the exceptional Lie groups.</p> Answer 6: <p>There is actually a collection of more general statements that detail the structure of the $\pi_4$.</p> <p>Firstly it was Browder [The Cohomology of Covering Spaces of H-Spaces] who proved that a simply connected finite H-space is 2-connected. Clark [On $\pi_3$ of Finite Dimensional H-Spaces] showed that the first non-vanishing homotopy group of a finite loop space appeared in degrees 1 or 3 (the result that $\pi_2=0$ for a simply connected finite H-space also appears in Browders [Torsion in H-Spaces], see also Lin's [The First Homotopy Group of a Finite H-Space]). Kane and Hubbuck [On $\pi_3$ of a Finite H-Space] prove that $\pi_3$ is torsion free, building on previous work of Thomas and Lin which required stricter assumptions (the Lie group $E_8$ does not fall under Thomas's requirements, for instance).</p> <p>Finally Hubbuck [Homotopy Groups of Finite H-Spaces] studied $\pi_4$, showing that it was a direct sum of finite groups of order 2. The proof uses JHC Whitehead's Certain Exact Sequence (see Baues [Homotopy Type and Homology] for a good discussion), identifying the groups that appear within using the work of others. The end result is an exact sequence</p> <p>$H_5(X;\mathbb{Z})\xrightarrow{\rho_2} H_5(X;\mathbb{Z}_2)\xrightarrow{Sq^2_*}H_3(X;\mathbb{Z}_2)\rightarrow \pi_4X\rightarrow0$</p> <p>And we know that $H_3(X;\mathbb{Z}_2)=\oplus\,\mathbb{Z}_2$ from the Hurewicz theorem and the work of Kane and Hubbuck so the result on $\pi_4X$ follows.</p> <p>Harper claims that if $X$ is simply connected and $H_*(\Omega X;\mathbb{Z})$ is torsion free then the result follows from the previous work of Bott and Samelson [Applications of the Theory of Morse to Symmetric Spaces].</p> Answer 7: <p>As a partial answer, here's at least a uniform statement, which can be found as Theorem 3.10 in Mimura's survey "Homotopy theory of Lie groups" in the Handbook of Algebraic Topology: </p> <blockquote> <p>Let $G$ be a compact, connected, simply connected, simple Lie group, $T$ a maximal torus of $G$, $\mathbb{R}^r$ its universal covering and $\Gamma$ the inverse image of the identity in $\mathbb{R}^r$. Let $a$ be the dominant root with respect to some lexicographic order of the roots of $G$. Then $\pi_4(G)$ is isomorphic to $\mathbb{Z}/2\mathbb{Z}$ if the hyperplane $a=1$ contains a point of $\Gamma$ and $0$ otherwise. </p> </blockquote> <p>(In the survey article, the statement has a $\mathbb{Z}$ where I wrote $\mathbb{Z}/2\mathbb{Z}$ but that would appear to be a typo.)</p> <p>Mimura refers to the paper</p> <ul> <li>R. Bott and H. Samelson. Applications of the theory of Morse to symmetric spaces. Amer. J. Math. 80 (1958), 964-1029. </li> </ul> <p>I can't check right now, but the uniform statement suggests a uniform proof, relating the fourth homotopy group to information on the root system. It would seem that the difference that singles out the symplectic groups is that the long roots are divisible (by 2 incidentally) in the character group.</p> Answer 8: <p>Since you are asking a question about $\pi_4$, lets use a little homotopy theory to think about this, and for starters, we can just ponder compact Lie groups. </p> <p>Consider first $G=Sp(1)=SU(2)=S^3$. It is well known that $\pi_4(S^3) = \mathbb Z/2$. </p> <p>Now lets compare and contrast $SU(3)$ with $Sp(2)$. $SU(3)/SU(2) = S^5$, while $Sp(2)/Sp(1) = S^7$, and these lead to fibration sequences $ S^3 \rightarrow SU(3) \rightarrow S^5$ and $ S^3 \rightarrow Sp(2) \rightarrow S^7$. The associated long exact sequences of homotopy groups make it obvious that $\pi_4(Sp(2)) = \mathbb Z/2$; more surprising from this perspective is that $\pi_4(SU(3)) = 0$.</p> <p>In general, well chosen (= highly connected) homogeneous spaces $G/H$ let one get at $\pi_*(G)$ from knowledge of $\pi_*(H)$ for $*$ not too large. For example, $G_2/SU(3) = S^6$, so $\pi_4(G_2) = 0$.</p> <p>Finally, here is a general method: Let $G\langle 3\rangle$ be the fiber of the fundamental class $G \rightarrow K(\mathbb Z,3)$. Then $\pi_4(G) = \pi_4(G\langle 3\rangle) = H_4(G\langle 3\rangle;\mathbb Z)$. Then one can try to get at this homology group using the Serre spectral sequence associated to the fibration $K(\mathbb Z,2) \rightarrow G\langle 3 \rangle \rightarrow G$. </p> <p>In various papers, M. Mimura has done much analysis of this sort, and in particular computes the first couple dozen homotopy groups of all of the exceptional Lie groups.</p> Answer 9: <p>There is actually a collection of more general statements that detail the structure of the $\pi_4$.</p> <p>Firstly it was Browder [The Cohomology of Covering Spaces of H-Spaces] who proved that a simply connected finite H-space is 2-connected. Clark [On $\pi_3$ of Finite Dimensional H-Spaces] showed that the first non-vanishing homotopy group of a finite loop space appeared in degrees 1 or 3 (the result that $\pi_2=0$ for a simply connected finite H-space also appears in Browders [Torsion in H-Spaces], see also Lin's [The First Homotopy Group of a Finite H-Space]). Kane and Hubbuck [On $\pi_3$ of a Finite H-Space] prove that $\pi_3$ is torsion free, building on previous work of Thomas and Lin which required stricter assumptions (the Lie group $E_8$ does not fall under Thomas's requirements, for instance).</p> <p>Finally Hubbuck [Homotopy Groups of Finite H-Spaces] studied $\pi_4$, showing that it was a direct sum of finite groups of order 2. The proof uses JHC Whitehead's Certain Exact Sequence (see Baues [Homotopy Type and Homology] for a good discussion), identifying the groups that appear within using the work of others. The end result is an exact sequence</p> <p>$H_5(X;\mathbb{Z})\xrightarrow{\rho_2} H_5(X;\mathbb{Z}_2)\xrightarrow{Sq^2_*}H_3(X;\mathbb{Z}_2)\rightarrow \pi_4X\rightarrow0$</p> <p>And we know that $H_3(X;\mathbb{Z}_2)=\oplus\,\mathbb{Z}_2$ from the Hurewicz theorem and the work of Kane and Hubbuck so the result on $\pi_4X$ follows.</p> <p>Harper claims that if $X$ is simply connected and $H_*(\Omega X;\mathbb{Z})$ is torsion free then the result follows from the previous work of Bott and Samelson [Applications of the Theory of Morse to Symmetric Spaces].</p>
Has the Jacobson/ Baer radical of a group been studied?
259,490
13
https://mathoverflow.net/questions/259490/has-the-jacobson-baer-radical-of-a-group-been-studied
<p>On <a href="https://groupprops.subwiki.org/wiki/Jacobson_radical">groupprops</a>, the Jacobson or Baer radical of a group $G$ is defined to be the intersection of all maximal normal subgroups of $G$. This is similar to, but distinct from, the Frattini subgroup which is the intersection of all maximal subgroups of $G$. </p> <p>Obviously the Jacobson radical of rings and the Frattini subgroup have been well-studied and are very useful.</p> <p>Are there any significant results concerning or uses of the Jacobson radical of groups?</p> <p>For comparison, significant results for the Frattini subgroup would include the facts about in on the <a href="https://en.wikipedia.org/wiki/Frattini_subgroup">Wikipedia page</a>. So equivalent characterisations, its properties and so on.</p>
[ { "answer_id": 259545, "score": 18, "text": "<p>The Jacobson radical $\\mathfrak{J}(G)$ of a group $G$ is defined by Reinhold Baer in [2] as the intersection of the maximal normal subgroups of $G$ (as noted by Baer the identity $G = \\mathfrak{J}(G)$ holds if and only if $G$ has no maximal normal subgroup). His seminal article [2] was written in German and doesn't seem to be widely known. It contains plenty of interesting elementary results, some of which have been reproved multiple times by different authors. Here is a non-comprehensive list.</p>\n\n<ul>\n<li>$\\mathfrak{J}(G/\\mathfrak{J}(G)) = 1$.</li>\n<li>$\\mathfrak{J}(G)$ is the set of all the elements which can be omitted from any set which generates $G$ as a normal subgroup, i.e., $\\mathfrak{J}(G)$ is the set of elements $g \\in G$ such for every $X \\subset G$, $\\langle X, g \\rangle^G = G$ implies $\\langle X \\rangle^G = G$ [2, Satz 3.1.d].</li>\n<li>For any two groups $A$ and $B$, we have $\\mathfrak{J}(A \\times B) = \\mathfrak{J}(A) \\times \\mathfrak{J}(B)$ [2, Folgerung 3.4].</li>\n<li>If $G$ is soluble then $[G, G] \\subset \\mathfrak{J}(G)$ [2, Folgerung 2.10].</li>\n<li>Assume that $\\mathfrak{J}(G) \\neq G$ and that the set of normal subgroups of $G/\\mathfrak{J}(G)$ which are contained in $[G/\\mathfrak{J}(G), G/\\mathfrak{J}(G)]$ satisfies the minimal condition. Then a subset generates $G$ as a normal subgroup if and only if its image generates $G/[G, G]$ [2, Satz 6.4], see [5] for a generalization.</li>\n</ul>\n\n<p>Of course, $\\mathfrak{J}(G)$ is a characteristic subgroup of $G$ and it is not difficult to show that $\\mathfrak{J}(G)$ contains the Frattini subgroup of $G$, <strong>provided that every simple quotient of $G$ has a maximal subgroup (see Corrigendum below)</strong>.</p>\n\n<p>I discovered the definition of the Jacobson/Baer radical $\\mathfrak{J}(G)$ and some of its elementary properties when studying two generalizations of the Andrews-Curtis conjecture. The authors of [7, 8] refer to $\\mathfrak{J}(G)$ as the $N$-Frattini subgroup of $G$ and denote it by $W(G)$. The letter $W$ primary alludes to the word \"weight\", i.e., the minimal number $w(G)$ of normal generators as commonly defined by knot theoricists, see <a href=\"https://mathoverflow.net/questions/54964/how-many-elements-does-it-take-to-normally-generate-a-group\">this MO post</a> for more about $w(G)$. But it could also be a(n) (un)conscious tribute to James Wiegold. \nIn [5], the concept of Jacobson radical is further generalized to $\\Omega$-groups for which the first two items above still hold, as well as some other properties [7, Lemma 4.1]. To echo Geoff Robinson's comment, it is proved in [7, Lemma 4.5] that a finite group $G$ is perfect if and only if $G/\\mathfrak{J}(G)$ is the direct product of simple non-Abelian groups. The use of $\\mathfrak{J}(G)$ is instrumental in the proof of the finitary Andrews-Curtis conjecture [7] and its quantitative enhancements [8].</p>\n\n<p>In a preprint of mine, I made this humble observation about finitely generated groups:\nAny two $n$-tuples which generate $G$ as a normal subgroup can be related by a finite sequence of Andrews-Curtis moves if the two following hold.</p>\n\n<ul>\n<li>$n &gt; w(G)$; </li>\n<li>$G/\\mathfrak{J}(G)$ is Abelian, or finite, or a direct product of finitely many non-Abelian simple groups.</li>\n</ul>\n\n<p><strong>Corrigendum:</strong>\nLet $G$ be any group and let $M$ be a maximal normal subgroup of $G$.\nLemma 4 of [1]\n(<a href=\"http://www.ams.org/mathscinet-getitem?mr=0122885\" rel=\"noreferrer\">MR0122885</a>) asserts that if $M$ is contained in some maximal subgroup of $G$, then $M$ contains the Frattini subgroup $\\Phi(G)$ of $G$. As a result, $\\Phi(G) \\subset \\mathfrak{J}(G)$ holds if every simple quotient of $G$ has a maximal subgroup, e.g., $G$ is finitely generated or $G$ is soluble.</p>\n\n<p>The existence or non-existence of a simple group without maximal subgroups was an unsolved problem in 1970 when B. H. Neumann reviewed [1]. It was still an open problem in 1974 as the analogous problem for near simple group was unsolved at that time, see [3] (<a href=\"http://www.ams.org/mathscinet-getitem?mr=0347986\" rel=\"noreferrer\">MR0347986</a>).</p>\n\n<p>The existence of an uncountable simple group without maximal subgroups was eventually established by Shelah in 1980 [4]. The existence of a countable simple group without maximal subgroups was established by Ol′shanskiĭ in 1991 [6, Theorem 35.3]. If $G$ is any of these two infinite simple groups, then we have $\\Phi(G) = G$ and $\\mathfrak{J}(G) = 1$ so that $\\Phi(G) \\not\\subset \\mathfrak{J}(G)$.</p>\n\n<hr>\n\n<p>[1] V. Dlab and V. Kořínek, \"The Frattini subgroup of a direct product of groups\", 1960.<br>\n[2] R. Baer, \"Der reduzierte Rang einer Gruppe\", 1964.<br>\n[3] J. Riles, \"The near Frattini subgroups of direct products of groups\", 1974.<br>\n[4] S. Shelah, \"On a problem of Kurosh, Jónsson groups, and applications\", 1980.<br>\n[5] A. Rhemtulla, \"Groups of finite weight\", 1981.<br>\n[6] A. Ol′shanskiĭ, \"Geometry of defining relations in groups\", 1991.<br>\n[7] A. Borovik, A. Lubotzky, and A. Myasnikov, \"The finitary Andrews-Curtis conjecture\", 2005.<br>\n[8] R. Burns and D. Oancea, \"Recalcitrance in groups II\", 2012.</p>\n" }, { "answer_id": 259561, "score": 4, "text": "<p>In addition to applications to finite/finitely generated groups, it's also useful in the study of locally compact groups (taking a definition that only allows closed normal subgroups). Other names for it include the cosocle, the Mel'nikov subgroup or the normal Frattini subgroup.</p>\n\n<p>In a connected Lie group $G$, $J(G)$ is the largest (not necessarily connected) soluble closed normal subgroup of $G$, and $G/J(G)$ is a direct product of simple Lie groups. Given the Gleason-Yamabe theorem, similar statements apply more generally to connected locally compact groups.</p>\n\n<p>If $G$ is a profinite group, then every proper closed normal subgroup is contained in a maximal open normal subgroup, and $G/J(G)$ is the largest quotient that is a Cartesian product of finite simple groups. Moreover, the series $G &gt; J(G) &gt; J(J(G)) &gt; \\dots$ always has trivial intersection, so it can be a useful characteristic series, especially if the terms are known to have finite index (equivalently, if $J(H)$ has finite index for every open subgroup $H$). Zalesskii showed that if $G$ is an infinite profinite group and $J(H)$ has finite index for every open subgroup $H$, then $G$ has a just infinite quotient.</p>\n\n<p>If $G$ is a compactly generated locally compact group with no infinite discrete quotient, then Caprace--Monod showed that there is a smallest cocompact normal subgroup $N$ of $G$, and $N/J(N)$ is generated topologically by a finite set of topologically simple closed normal subgroups, with $N &gt; J(N)$ unless $N$ is connected and soluble.</p>\n" }, { "answer_id": 259545, "score": 18, "text": "<p>The Jacobson radical $\\mathfrak{J}(G)$ of a group $G$ is defined by Reinhold Baer in [2] as the intersection of the maximal normal subgroups of $G$ (as noted by Baer the identity $G = \\mathfrak{J}(G)$ holds if and only if $G$ has no maximal normal subgroup). His seminal article [2] was written in German and doesn't seem to be widely known. It contains plenty of interesting elementary results, some of which have been reproved multiple times by different authors. Here is a non-comprehensive list.</p>\n\n<ul>\n<li>$\\mathfrak{J}(G/\\mathfrak{J}(G)) = 1$.</li>\n<li>$\\mathfrak{J}(G)$ is the set of all the elements which can be omitted from any set which generates $G$ as a normal subgroup, i.e., $\\mathfrak{J}(G)$ is the set of elements $g \\in G$ such for every $X \\subset G$, $\\langle X, g \\rangle^G = G$ implies $\\langle X \\rangle^G = G$ [2, Satz 3.1.d].</li>\n<li>For any two groups $A$ and $B$, we have $\\mathfrak{J}(A \\times B) = \\mathfrak{J}(A) \\times \\mathfrak{J}(B)$ [2, Folgerung 3.4].</li>\n<li>If $G$ is soluble then $[G, G] \\subset \\mathfrak{J}(G)$ [2, Folgerung 2.10].</li>\n<li>Assume that $\\mathfrak{J}(G) \\neq G$ and that the set of normal subgroups of $G/\\mathfrak{J}(G)$ which are contained in $[G/\\mathfrak{J}(G), G/\\mathfrak{J}(G)]$ satisfies the minimal condition. Then a subset generates $G$ as a normal subgroup if and only if its image generates $G/[G, G]$ [2, Satz 6.4], see [5] for a generalization.</li>\n</ul>\n\n<p>Of course, $\\mathfrak{J}(G)$ is a characteristic subgroup of $G$ and it is not difficult to show that $\\mathfrak{J}(G)$ contains the Frattini subgroup of $G$, <strong>provided that every simple quotient of $G$ has a maximal subgroup (see Corrigendum below)</strong>.</p>\n\n<p>I discovered the definition of the Jacobson/Baer radical $\\mathfrak{J}(G)$ and some of its elementary properties when studying two generalizations of the Andrews-Curtis conjecture. The authors of [7, 8] refer to $\\mathfrak{J}(G)$ as the $N$-Frattini subgroup of $G$ and denote it by $W(G)$. The letter $W$ primary alludes to the word \"weight\", i.e., the minimal number $w(G)$ of normal generators as commonly defined by knot theoricists, see <a href=\"https://mathoverflow.net/questions/54964/how-many-elements-does-it-take-to-normally-generate-a-group\">this MO post</a> for more about $w(G)$. But it could also be a(n) (un)conscious tribute to James Wiegold. \nIn [5], the concept of Jacobson radical is further generalized to $\\Omega$-groups for which the first two items above still hold, as well as some other properties [7, Lemma 4.1]. To echo Geoff Robinson's comment, it is proved in [7, Lemma 4.5] that a finite group $G$ is perfect if and only if $G/\\mathfrak{J}(G)$ is the direct product of simple non-Abelian groups. The use of $\\mathfrak{J}(G)$ is instrumental in the proof of the finitary Andrews-Curtis conjecture [7] and its quantitative enhancements [8].</p>\n\n<p>In a preprint of mine, I made this humble observation about finitely generated groups:\nAny two $n$-tuples which generate $G$ as a normal subgroup can be related by a finite sequence of Andrews-Curtis moves if the two following hold.</p>\n\n<ul>\n<li>$n &gt; w(G)$; </li>\n<li>$G/\\mathfrak{J}(G)$ is Abelian, or finite, or a direct product of finitely many non-Abelian simple groups.</li>\n</ul>\n\n<p><strong>Corrigendum:</strong>\nLet $G$ be any group and let $M$ be a maximal normal subgroup of $G$.\nLemma 4 of [1]\n(<a href=\"http://www.ams.org/mathscinet-getitem?mr=0122885\" rel=\"noreferrer\">MR0122885</a>) asserts that if $M$ is contained in some maximal subgroup of $G$, then $M$ contains the Frattini subgroup $\\Phi(G)$ of $G$. As a result, $\\Phi(G) \\subset \\mathfrak{J}(G)$ holds if every simple quotient of $G$ has a maximal subgroup, e.g., $G$ is finitely generated or $G$ is soluble.</p>\n\n<p>The existence or non-existence of a simple group without maximal subgroups was an unsolved problem in 1970 when B. H. Neumann reviewed [1]. It was still an open problem in 1974 as the analogous problem for near simple group was unsolved at that time, see [3] (<a href=\"http://www.ams.org/mathscinet-getitem?mr=0347986\" rel=\"noreferrer\">MR0347986</a>).</p>\n\n<p>The existence of an uncountable simple group without maximal subgroups was eventually established by Shelah in 1980 [4]. The existence of a countable simple group without maximal subgroups was established by Ol′shanskiĭ in 1991 [6, Theorem 35.3]. If $G$ is any of these two infinite simple groups, then we have $\\Phi(G) = G$ and $\\mathfrak{J}(G) = 1$ so that $\\Phi(G) \\not\\subset \\mathfrak{J}(G)$.</p>\n\n<hr>\n\n<p>[1] V. Dlab and V. Kořínek, \"The Frattini subgroup of a direct product of groups\", 1960.<br>\n[2] R. Baer, \"Der reduzierte Rang einer Gruppe\", 1964.<br>\n[3] J. Riles, \"The near Frattini subgroups of direct products of groups\", 1974.<br>\n[4] S. Shelah, \"On a problem of Kurosh, Jónsson groups, and applications\", 1980.<br>\n[5] A. Rhemtulla, \"Groups of finite weight\", 1981.<br>\n[6] A. Ol′shanskiĭ, \"Geometry of defining relations in groups\", 1991.<br>\n[7] A. Borovik, A. Lubotzky, and A. Myasnikov, \"The finitary Andrews-Curtis conjecture\", 2005.<br>\n[8] R. Burns and D. Oancea, \"Recalcitrance in groups II\", 2012.</p>\n" }, { "answer_id": 259561, "score": 4, "text": "<p>In addition to applications to finite/finitely generated groups, it's also useful in the study of locally compact groups (taking a definition that only allows closed normal subgroups). Other names for it include the cosocle, the Mel'nikov subgroup or the normal Frattini subgroup.</p>\n\n<p>In a connected Lie group $G$, $J(G)$ is the largest (not necessarily connected) soluble closed normal subgroup of $G$, and $G/J(G)$ is a direct product of simple Lie groups. Given the Gleason-Yamabe theorem, similar statements apply more generally to connected locally compact groups.</p>\n\n<p>If $G$ is a profinite group, then every proper closed normal subgroup is contained in a maximal open normal subgroup, and $G/J(G)$ is the largest quotient that is a Cartesian product of finite simple groups. Moreover, the series $G &gt; J(G) &gt; J(J(G)) &gt; \\dots$ always has trivial intersection, so it can be a useful characteristic series, especially if the terms are known to have finite index (equivalently, if $J(H)$ has finite index for every open subgroup $H$). Zalesskii showed that if $G$ is an infinite profinite group and $J(H)$ has finite index for every open subgroup $H$, then $G$ has a just infinite quotient.</p>\n\n<p>If $G$ is a compactly generated locally compact group with no infinite discrete quotient, then Caprace--Monod showed that there is a smallest cocompact normal subgroup $N$ of $G$, and $N/J(N)$ is generated topologically by a finite set of topologically simple closed normal subgroups, with $N &gt; J(N)$ unless $N$ is connected and soluble.</p>\n" }, { "answer_id": 259545, "score": 18, "text": "<p>The Jacobson radical $\\mathfrak{J}(G)$ of a group $G$ is defined by Reinhold Baer in [2] as the intersection of the maximal normal subgroups of $G$ (as noted by Baer the identity $G = \\mathfrak{J}(G)$ holds if and only if $G$ has no maximal normal subgroup). His seminal article [2] was written in German and doesn't seem to be widely known. It contains plenty of interesting elementary results, some of which have been reproved multiple times by different authors. Here is a non-comprehensive list.</p>\n\n<ul>\n<li>$\\mathfrak{J}(G/\\mathfrak{J}(G)) = 1$.</li>\n<li>$\\mathfrak{J}(G)$ is the set of all the elements which can be omitted from any set which generates $G$ as a normal subgroup, i.e., $\\mathfrak{J}(G)$ is the set of elements $g \\in G$ such for every $X \\subset G$, $\\langle X, g \\rangle^G = G$ implies $\\langle X \\rangle^G = G$ [2, Satz 3.1.d].</li>\n<li>For any two groups $A$ and $B$, we have $\\mathfrak{J}(A \\times B) = \\mathfrak{J}(A) \\times \\mathfrak{J}(B)$ [2, Folgerung 3.4].</li>\n<li>If $G$ is soluble then $[G, G] \\subset \\mathfrak{J}(G)$ [2, Folgerung 2.10].</li>\n<li>Assume that $\\mathfrak{J}(G) \\neq G$ and that the set of normal subgroups of $G/\\mathfrak{J}(G)$ which are contained in $[G/\\mathfrak{J}(G), G/\\mathfrak{J}(G)]$ satisfies the minimal condition. Then a subset generates $G$ as a normal subgroup if and only if its image generates $G/[G, G]$ [2, Satz 6.4], see [5] for a generalization.</li>\n</ul>\n\n<p>Of course, $\\mathfrak{J}(G)$ is a characteristic subgroup of $G$ and it is not difficult to show that $\\mathfrak{J}(G)$ contains the Frattini subgroup of $G$, <strong>provided that every simple quotient of $G$ has a maximal subgroup (see Corrigendum below)</strong>.</p>\n\n<p>I discovered the definition of the Jacobson/Baer radical $\\mathfrak{J}(G)$ and some of its elementary properties when studying two generalizations of the Andrews-Curtis conjecture. The authors of [7, 8] refer to $\\mathfrak{J}(G)$ as the $N$-Frattini subgroup of $G$ and denote it by $W(G)$. The letter $W$ primary alludes to the word \"weight\", i.e., the minimal number $w(G)$ of normal generators as commonly defined by knot theoricists, see <a href=\"https://mathoverflow.net/questions/54964/how-many-elements-does-it-take-to-normally-generate-a-group\">this MO post</a> for more about $w(G)$. But it could also be a(n) (un)conscious tribute to James Wiegold. \nIn [5], the concept of Jacobson radical is further generalized to $\\Omega$-groups for which the first two items above still hold, as well as some other properties [7, Lemma 4.1]. To echo Geoff Robinson's comment, it is proved in [7, Lemma 4.5] that a finite group $G$ is perfect if and only if $G/\\mathfrak{J}(G)$ is the direct product of simple non-Abelian groups. The use of $\\mathfrak{J}(G)$ is instrumental in the proof of the finitary Andrews-Curtis conjecture [7] and its quantitative enhancements [8].</p>\n\n<p>In a preprint of mine, I made this humble observation about finitely generated groups:\nAny two $n$-tuples which generate $G$ as a normal subgroup can be related by a finite sequence of Andrews-Curtis moves if the two following hold.</p>\n\n<ul>\n<li>$n &gt; w(G)$; </li>\n<li>$G/\\mathfrak{J}(G)$ is Abelian, or finite, or a direct product of finitely many non-Abelian simple groups.</li>\n</ul>\n\n<p><strong>Corrigendum:</strong>\nLet $G$ be any group and let $M$ be a maximal normal subgroup of $G$.\nLemma 4 of [1]\n(<a href=\"http://www.ams.org/mathscinet-getitem?mr=0122885\" rel=\"noreferrer\">MR0122885</a>) asserts that if $M$ is contained in some maximal subgroup of $G$, then $M$ contains the Frattini subgroup $\\Phi(G)$ of $G$. As a result, $\\Phi(G) \\subset \\mathfrak{J}(G)$ holds if every simple quotient of $G$ has a maximal subgroup, e.g., $G$ is finitely generated or $G$ is soluble.</p>\n\n<p>The existence or non-existence of a simple group without maximal subgroups was an unsolved problem in 1970 when B. H. Neumann reviewed [1]. It was still an open problem in 1974 as the analogous problem for near simple group was unsolved at that time, see [3] (<a href=\"http://www.ams.org/mathscinet-getitem?mr=0347986\" rel=\"noreferrer\">MR0347986</a>).</p>\n\n<p>The existence of an uncountable simple group without maximal subgroups was eventually established by Shelah in 1980 [4]. The existence of a countable simple group without maximal subgroups was established by Ol′shanskiĭ in 1991 [6, Theorem 35.3]. If $G$ is any of these two infinite simple groups, then we have $\\Phi(G) = G$ and $\\mathfrak{J}(G) = 1$ so that $\\Phi(G) \\not\\subset \\mathfrak{J}(G)$.</p>\n\n<hr>\n\n<p>[1] V. Dlab and V. Kořínek, \"The Frattini subgroup of a direct product of groups\", 1960.<br>\n[2] R. Baer, \"Der reduzierte Rang einer Gruppe\", 1964.<br>\n[3] J. Riles, \"The near Frattini subgroups of direct products of groups\", 1974.<br>\n[4] S. Shelah, \"On a problem of Kurosh, Jónsson groups, and applications\", 1980.<br>\n[5] A. Rhemtulla, \"Groups of finite weight\", 1981.<br>\n[6] A. Ol′shanskiĭ, \"Geometry of defining relations in groups\", 1991.<br>\n[7] A. Borovik, A. Lubotzky, and A. Myasnikov, \"The finitary Andrews-Curtis conjecture\", 2005.<br>\n[8] R. Burns and D. Oancea, \"Recalcitrance in groups II\", 2012.</p>\n" }, { "answer_id": 259561, "score": 4, "text": "<p>In addition to applications to finite/finitely generated groups, it's also useful in the study of locally compact groups (taking a definition that only allows closed normal subgroups). Other names for it include the cosocle, the Mel'nikov subgroup or the normal Frattini subgroup.</p>\n\n<p>In a connected Lie group $G$, $J(G)$ is the largest (not necessarily connected) soluble closed normal subgroup of $G$, and $G/J(G)$ is a direct product of simple Lie groups. Given the Gleason-Yamabe theorem, similar statements apply more generally to connected locally compact groups.</p>\n\n<p>If $G$ is a profinite group, then every proper closed normal subgroup is contained in a maximal open normal subgroup, and $G/J(G)$ is the largest quotient that is a Cartesian product of finite simple groups. Moreover, the series $G &gt; J(G) &gt; J(J(G)) &gt; \\dots$ always has trivial intersection, so it can be a useful characteristic series, especially if the terms are known to have finite index (equivalently, if $J(H)$ has finite index for every open subgroup $H$). Zalesskii showed that if $G$ is an infinite profinite group and $J(H)$ has finite index for every open subgroup $H$, then $G$ has a just infinite quotient.</p>\n\n<p>If $G$ is a compactly generated locally compact group with no infinite discrete quotient, then Caprace--Monod showed that there is a smallest cocompact normal subgroup $N$ of $G$, and $N/J(N)$ is generated topologically by a finite set of topologically simple closed normal subgroups, with $N &gt; J(N)$ unless $N$ is connected and soluble.</p>\n" } ]
Question: Has the Jacobson/ Baer radical of a group been studied? Answer 1: <p>The Jacobson radical $\mathfrak{J}(G)$ of a group $G$ is defined by Reinhold Baer in [2] as the intersection of the maximal normal subgroups of $G$ (as noted by Baer the identity $G = \mathfrak{J}(G)$ holds if and only if $G$ has no maximal normal subgroup). His seminal article [2] was written in German and doesn't seem to be widely known. It contains plenty of interesting elementary results, some of which have been reproved multiple times by different authors. Here is a non-comprehensive list.</p> <ul> <li>$\mathfrak{J}(G/\mathfrak{J}(G)) = 1$.</li> <li>$\mathfrak{J}(G)$ is the set of all the elements which can be omitted from any set which generates $G$ as a normal subgroup, i.e., $\mathfrak{J}(G)$ is the set of elements $g \in G$ such for every $X \subset G$, $\langle X, g \rangle^G = G$ implies $\langle X \rangle^G = G$ [2, Satz 3.1.d].</li> <li>For any two groups $A$ and $B$, we have $\mathfrak{J}(A \times B) = \mathfrak{J}(A) \times \mathfrak{J}(B)$ [2, Folgerung 3.4].</li> <li>If $G$ is soluble then $[G, G] \subset \mathfrak{J}(G)$ [2, Folgerung 2.10].</li> <li>Assume that $\mathfrak{J}(G) \neq G$ and that the set of normal subgroups of $G/\mathfrak{J}(G)$ which are contained in $[G/\mathfrak{J}(G), G/\mathfrak{J}(G)]$ satisfies the minimal condition. Then a subset generates $G$ as a normal subgroup if and only if its image generates $G/[G, G]$ [2, Satz 6.4], see [5] for a generalization.</li> </ul> <p>Of course, $\mathfrak{J}(G)$ is a characteristic subgroup of $G$ and it is not difficult to show that $\mathfrak{J}(G)$ contains the Frattini subgroup of $G$, <strong>provided that every simple quotient of $G$ has a maximal subgroup (see Corrigendum below)</strong>.</p> <p>I discovered the definition of the Jacobson/Baer radical $\mathfrak{J}(G)$ and some of its elementary properties when studying two generalizations of the Andrews-Curtis conjecture. The authors of [7, 8] refer to $\mathfrak{J}(G)$ as the $N$-Frattini subgroup of $G$ and denote it by $W(G)$. The letter $W$ primary alludes to the word "weight", i.e., the minimal number $w(G)$ of normal generators as commonly defined by knot theoricists, see <a href="https://mathoverflow.net/questions/54964/how-many-elements-does-it-take-to-normally-generate-a-group">this MO post</a> for more about $w(G)$. But it could also be a(n) (un)conscious tribute to James Wiegold. In [5], the concept of Jacobson radical is further generalized to $\Omega$-groups for which the first two items above still hold, as well as some other properties [7, Lemma 4.1]. To echo Geoff Robinson's comment, it is proved in [7, Lemma 4.5] that a finite group $G$ is perfect if and only if $G/\mathfrak{J}(G)$ is the direct product of simple non-Abelian groups. The use of $\mathfrak{J}(G)$ is instrumental in the proof of the finitary Andrews-Curtis conjecture [7] and its quantitative enhancements [8].</p> <p>In a preprint of mine, I made this humble observation about finitely generated groups: Any two $n$-tuples which generate $G$ as a normal subgroup can be related by a finite sequence of Andrews-Curtis moves if the two following hold.</p> <ul> <li>$n &gt; w(G)$; </li> <li>$G/\mathfrak{J}(G)$ is Abelian, or finite, or a direct product of finitely many non-Abelian simple groups.</li> </ul> <p><strong>Corrigendum:</strong> Let $G$ be any group and let $M$ be a maximal normal subgroup of $G$. Lemma 4 of [1] (<a href="http://www.ams.org/mathscinet-getitem?mr=0122885" rel="noreferrer">MR0122885</a>) asserts that if $M$ is contained in some maximal subgroup of $G$, then $M$ contains the Frattini subgroup $\Phi(G)$ of $G$. As a result, $\Phi(G) \subset \mathfrak{J}(G)$ holds if every simple quotient of $G$ has a maximal subgroup, e.g., $G$ is finitely generated or $G$ is soluble.</p> <p>The existence or non-existence of a simple group without maximal subgroups was an unsolved problem in 1970 when B. H. Neumann reviewed [1]. It was still an open problem in 1974 as the analogous problem for near simple group was unsolved at that time, see [3] (<a href="http://www.ams.org/mathscinet-getitem?mr=0347986" rel="noreferrer">MR0347986</a>).</p> <p>The existence of an uncountable simple group without maximal subgroups was eventually established by Shelah in 1980 [4]. The existence of a countable simple group without maximal subgroups was established by Ol′shanskiĭ in 1991 [6, Theorem 35.3]. If $G$ is any of these two infinite simple groups, then we have $\Phi(G) = G$ and $\mathfrak{J}(G) = 1$ so that $\Phi(G) \not\subset \mathfrak{J}(G)$.</p> <hr> <p>[1] V. Dlab and V. Kořínek, "The Frattini subgroup of a direct product of groups", 1960.<br> [2] R. Baer, "Der reduzierte Rang einer Gruppe", 1964.<br> [3] J. Riles, "The near Frattini subgroups of direct products of groups", 1974.<br> [4] S. Shelah, "On a problem of Kurosh, Jónsson groups, and applications", 1980.<br> [5] A. Rhemtulla, "Groups of finite weight", 1981.<br> [6] A. Ol′shanskiĭ, "Geometry of defining relations in groups", 1991.<br> [7] A. Borovik, A. Lubotzky, and A. Myasnikov, "The finitary Andrews-Curtis conjecture", 2005.<br> [8] R. Burns and D. Oancea, "Recalcitrance in groups II", 2012.</p> Answer 2: <p>In addition to applications to finite/finitely generated groups, it's also useful in the study of locally compact groups (taking a definition that only allows closed normal subgroups). Other names for it include the cosocle, the Mel'nikov subgroup or the normal Frattini subgroup.</p> <p>In a connected Lie group $G$, $J(G)$ is the largest (not necessarily connected) soluble closed normal subgroup of $G$, and $G/J(G)$ is a direct product of simple Lie groups. Given the Gleason-Yamabe theorem, similar statements apply more generally to connected locally compact groups.</p> <p>If $G$ is a profinite group, then every proper closed normal subgroup is contained in a maximal open normal subgroup, and $G/J(G)$ is the largest quotient that is a Cartesian product of finite simple groups. Moreover, the series $G &gt; J(G) &gt; J(J(G)) &gt; \dots$ always has trivial intersection, so it can be a useful characteristic series, especially if the terms are known to have finite index (equivalently, if $J(H)$ has finite index for every open subgroup $H$). Zalesskii showed that if $G$ is an infinite profinite group and $J(H)$ has finite index for every open subgroup $H$, then $G$ has a just infinite quotient.</p> <p>If $G$ is a compactly generated locally compact group with no infinite discrete quotient, then Caprace--Monod showed that there is a smallest cocompact normal subgroup $N$ of $G$, and $N/J(N)$ is generated topologically by a finite set of topologically simple closed normal subgroups, with $N &gt; J(N)$ unless $N$ is connected and soluble.</p> Answer 3: <p>The Jacobson radical $\mathfrak{J}(G)$ of a group $G$ is defined by Reinhold Baer in [2] as the intersection of the maximal normal subgroups of $G$ (as noted by Baer the identity $G = \mathfrak{J}(G)$ holds if and only if $G$ has no maximal normal subgroup). His seminal article [2] was written in German and doesn't seem to be widely known. It contains plenty of interesting elementary results, some of which have been reproved multiple times by different authors. Here is a non-comprehensive list.</p> <ul> <li>$\mathfrak{J}(G/\mathfrak{J}(G)) = 1$.</li> <li>$\mathfrak{J}(G)$ is the set of all the elements which can be omitted from any set which generates $G$ as a normal subgroup, i.e., $\mathfrak{J}(G)$ is the set of elements $g \in G$ such for every $X \subset G$, $\langle X, g \rangle^G = G$ implies $\langle X \rangle^G = G$ [2, Satz 3.1.d].</li> <li>For any two groups $A$ and $B$, we have $\mathfrak{J}(A \times B) = \mathfrak{J}(A) \times \mathfrak{J}(B)$ [2, Folgerung 3.4].</li> <li>If $G$ is soluble then $[G, G] \subset \mathfrak{J}(G)$ [2, Folgerung 2.10].</li> <li>Assume that $\mathfrak{J}(G) \neq G$ and that the set of normal subgroups of $G/\mathfrak{J}(G)$ which are contained in $[G/\mathfrak{J}(G), G/\mathfrak{J}(G)]$ satisfies the minimal condition. Then a subset generates $G$ as a normal subgroup if and only if its image generates $G/[G, G]$ [2, Satz 6.4], see [5] for a generalization.</li> </ul> <p>Of course, $\mathfrak{J}(G)$ is a characteristic subgroup of $G$ and it is not difficult to show that $\mathfrak{J}(G)$ contains the Frattini subgroup of $G$, <strong>provided that every simple quotient of $G$ has a maximal subgroup (see Corrigendum below)</strong>.</p> <p>I discovered the definition of the Jacobson/Baer radical $\mathfrak{J}(G)$ and some of its elementary properties when studying two generalizations of the Andrews-Curtis conjecture. The authors of [7, 8] refer to $\mathfrak{J}(G)$ as the $N$-Frattini subgroup of $G$ and denote it by $W(G)$. The letter $W$ primary alludes to the word "weight", i.e., the minimal number $w(G)$ of normal generators as commonly defined by knot theoricists, see <a href="https://mathoverflow.net/questions/54964/how-many-elements-does-it-take-to-normally-generate-a-group">this MO post</a> for more about $w(G)$. But it could also be a(n) (un)conscious tribute to James Wiegold. In [5], the concept of Jacobson radical is further generalized to $\Omega$-groups for which the first two items above still hold, as well as some other properties [7, Lemma 4.1]. To echo Geoff Robinson's comment, it is proved in [7, Lemma 4.5] that a finite group $G$ is perfect if and only if $G/\mathfrak{J}(G)$ is the direct product of simple non-Abelian groups. The use of $\mathfrak{J}(G)$ is instrumental in the proof of the finitary Andrews-Curtis conjecture [7] and its quantitative enhancements [8].</p> <p>In a preprint of mine, I made this humble observation about finitely generated groups: Any two $n$-tuples which generate $G$ as a normal subgroup can be related by a finite sequence of Andrews-Curtis moves if the two following hold.</p> <ul> <li>$n &gt; w(G)$; </li> <li>$G/\mathfrak{J}(G)$ is Abelian, or finite, or a direct product of finitely many non-Abelian simple groups.</li> </ul> <p><strong>Corrigendum:</strong> Let $G$ be any group and let $M$ be a maximal normal subgroup of $G$. Lemma 4 of [1] (<a href="http://www.ams.org/mathscinet-getitem?mr=0122885" rel="noreferrer">MR0122885</a>) asserts that if $M$ is contained in some maximal subgroup of $G$, then $M$ contains the Frattini subgroup $\Phi(G)$ of $G$. As a result, $\Phi(G) \subset \mathfrak{J}(G)$ holds if every simple quotient of $G$ has a maximal subgroup, e.g., $G$ is finitely generated or $G$ is soluble.</p> <p>The existence or non-existence of a simple group without maximal subgroups was an unsolved problem in 1970 when B. H. Neumann reviewed [1]. It was still an open problem in 1974 as the analogous problem for near simple group was unsolved at that time, see [3] (<a href="http://www.ams.org/mathscinet-getitem?mr=0347986" rel="noreferrer">MR0347986</a>).</p> <p>The existence of an uncountable simple group without maximal subgroups was eventually established by Shelah in 1980 [4]. The existence of a countable simple group without maximal subgroups was established by Ol′shanskiĭ in 1991 [6, Theorem 35.3]. If $G$ is any of these two infinite simple groups, then we have $\Phi(G) = G$ and $\mathfrak{J}(G) = 1$ so that $\Phi(G) \not\subset \mathfrak{J}(G)$.</p> <hr> <p>[1] V. Dlab and V. Kořínek, "The Frattini subgroup of a direct product of groups", 1960.<br> [2] R. Baer, "Der reduzierte Rang einer Gruppe", 1964.<br> [3] J. Riles, "The near Frattini subgroups of direct products of groups", 1974.<br> [4] S. Shelah, "On a problem of Kurosh, Jónsson groups, and applications", 1980.<br> [5] A. Rhemtulla, "Groups of finite weight", 1981.<br> [6] A. Ol′shanskiĭ, "Geometry of defining relations in groups", 1991.<br> [7] A. Borovik, A. Lubotzky, and A. Myasnikov, "The finitary Andrews-Curtis conjecture", 2005.<br> [8] R. Burns and D. Oancea, "Recalcitrance in groups II", 2012.</p> Answer 4: <p>In addition to applications to finite/finitely generated groups, it's also useful in the study of locally compact groups (taking a definition that only allows closed normal subgroups). Other names for it include the cosocle, the Mel'nikov subgroup or the normal Frattini subgroup.</p> <p>In a connected Lie group $G$, $J(G)$ is the largest (not necessarily connected) soluble closed normal subgroup of $G$, and $G/J(G)$ is a direct product of simple Lie groups. Given the Gleason-Yamabe theorem, similar statements apply more generally to connected locally compact groups.</p> <p>If $G$ is a profinite group, then every proper closed normal subgroup is contained in a maximal open normal subgroup, and $G/J(G)$ is the largest quotient that is a Cartesian product of finite simple groups. Moreover, the series $G &gt; J(G) &gt; J(J(G)) &gt; \dots$ always has trivial intersection, so it can be a useful characteristic series, especially if the terms are known to have finite index (equivalently, if $J(H)$ has finite index for every open subgroup $H$). Zalesskii showed that if $G$ is an infinite profinite group and $J(H)$ has finite index for every open subgroup $H$, then $G$ has a just infinite quotient.</p> <p>If $G$ is a compactly generated locally compact group with no infinite discrete quotient, then Caprace--Monod showed that there is a smallest cocompact normal subgroup $N$ of $G$, and $N/J(N)$ is generated topologically by a finite set of topologically simple closed normal subgroups, with $N &gt; J(N)$ unless $N$ is connected and soluble.</p> Answer 5: <p>The Jacobson radical $\mathfrak{J}(G)$ of a group $G$ is defined by Reinhold Baer in [2] as the intersection of the maximal normal subgroups of $G$ (as noted by Baer the identity $G = \mathfrak{J}(G)$ holds if and only if $G$ has no maximal normal subgroup). His seminal article [2] was written in German and doesn't seem to be widely known. It contains plenty of interesting elementary results, some of which have been reproved multiple times by different authors. Here is a non-comprehensive list.</p> <ul> <li>$\mathfrak{J}(G/\mathfrak{J}(G)) = 1$.</li> <li>$\mathfrak{J}(G)$ is the set of all the elements which can be omitted from any set which generates $G$ as a normal subgroup, i.e., $\mathfrak{J}(G)$ is the set of elements $g \in G$ such for every $X \subset G$, $\langle X, g \rangle^G = G$ implies $\langle X \rangle^G = G$ [2, Satz 3.1.d].</li> <li>For any two groups $A$ and $B$, we have $\mathfrak{J}(A \times B) = \mathfrak{J}(A) \times \mathfrak{J}(B)$ [2, Folgerung 3.4].</li> <li>If $G$ is soluble then $[G, G] \subset \mathfrak{J}(G)$ [2, Folgerung 2.10].</li> <li>Assume that $\mathfrak{J}(G) \neq G$ and that the set of normal subgroups of $G/\mathfrak{J}(G)$ which are contained in $[G/\mathfrak{J}(G), G/\mathfrak{J}(G)]$ satisfies the minimal condition. Then a subset generates $G$ as a normal subgroup if and only if its image generates $G/[G, G]$ [2, Satz 6.4], see [5] for a generalization.</li> </ul> <p>Of course, $\mathfrak{J}(G)$ is a characteristic subgroup of $G$ and it is not difficult to show that $\mathfrak{J}(G)$ contains the Frattini subgroup of $G$, <strong>provided that every simple quotient of $G$ has a maximal subgroup (see Corrigendum below)</strong>.</p> <p>I discovered the definition of the Jacobson/Baer radical $\mathfrak{J}(G)$ and some of its elementary properties when studying two generalizations of the Andrews-Curtis conjecture. The authors of [7, 8] refer to $\mathfrak{J}(G)$ as the $N$-Frattini subgroup of $G$ and denote it by $W(G)$. The letter $W$ primary alludes to the word "weight", i.e., the minimal number $w(G)$ of normal generators as commonly defined by knot theoricists, see <a href="https://mathoverflow.net/questions/54964/how-many-elements-does-it-take-to-normally-generate-a-group">this MO post</a> for more about $w(G)$. But it could also be a(n) (un)conscious tribute to James Wiegold. In [5], the concept of Jacobson radical is further generalized to $\Omega$-groups for which the first two items above still hold, as well as some other properties [7, Lemma 4.1]. To echo Geoff Robinson's comment, it is proved in [7, Lemma 4.5] that a finite group $G$ is perfect if and only if $G/\mathfrak{J}(G)$ is the direct product of simple non-Abelian groups. The use of $\mathfrak{J}(G)$ is instrumental in the proof of the finitary Andrews-Curtis conjecture [7] and its quantitative enhancements [8].</p> <p>In a preprint of mine, I made this humble observation about finitely generated groups: Any two $n$-tuples which generate $G$ as a normal subgroup can be related by a finite sequence of Andrews-Curtis moves if the two following hold.</p> <ul> <li>$n &gt; w(G)$; </li> <li>$G/\mathfrak{J}(G)$ is Abelian, or finite, or a direct product of finitely many non-Abelian simple groups.</li> </ul> <p><strong>Corrigendum:</strong> Let $G$ be any group and let $M$ be a maximal normal subgroup of $G$. Lemma 4 of [1] (<a href="http://www.ams.org/mathscinet-getitem?mr=0122885" rel="noreferrer">MR0122885</a>) asserts that if $M$ is contained in some maximal subgroup of $G$, then $M$ contains the Frattini subgroup $\Phi(G)$ of $G$. As a result, $\Phi(G) \subset \mathfrak{J}(G)$ holds if every simple quotient of $G$ has a maximal subgroup, e.g., $G$ is finitely generated or $G$ is soluble.</p> <p>The existence or non-existence of a simple group without maximal subgroups was an unsolved problem in 1970 when B. H. Neumann reviewed [1]. It was still an open problem in 1974 as the analogous problem for near simple group was unsolved at that time, see [3] (<a href="http://www.ams.org/mathscinet-getitem?mr=0347986" rel="noreferrer">MR0347986</a>).</p> <p>The existence of an uncountable simple group without maximal subgroups was eventually established by Shelah in 1980 [4]. The existence of a countable simple group without maximal subgroups was established by Ol′shanskiĭ in 1991 [6, Theorem 35.3]. If $G$ is any of these two infinite simple groups, then we have $\Phi(G) = G$ and $\mathfrak{J}(G) = 1$ so that $\Phi(G) \not\subset \mathfrak{J}(G)$.</p> <hr> <p>[1] V. Dlab and V. Kořínek, "The Frattini subgroup of a direct product of groups", 1960.<br> [2] R. Baer, "Der reduzierte Rang einer Gruppe", 1964.<br> [3] J. Riles, "The near Frattini subgroups of direct products of groups", 1974.<br> [4] S. Shelah, "On a problem of Kurosh, Jónsson groups, and applications", 1980.<br> [5] A. Rhemtulla, "Groups of finite weight", 1981.<br> [6] A. Ol′shanskiĭ, "Geometry of defining relations in groups", 1991.<br> [7] A. Borovik, A. Lubotzky, and A. Myasnikov, "The finitary Andrews-Curtis conjecture", 2005.<br> [8] R. Burns and D. Oancea, "Recalcitrance in groups II", 2012.</p> Answer 6: <p>In addition to applications to finite/finitely generated groups, it's also useful in the study of locally compact groups (taking a definition that only allows closed normal subgroups). Other names for it include the cosocle, the Mel'nikov subgroup or the normal Frattini subgroup.</p> <p>In a connected Lie group $G$, $J(G)$ is the largest (not necessarily connected) soluble closed normal subgroup of $G$, and $G/J(G)$ is a direct product of simple Lie groups. Given the Gleason-Yamabe theorem, similar statements apply more generally to connected locally compact groups.</p> <p>If $G$ is a profinite group, then every proper closed normal subgroup is contained in a maximal open normal subgroup, and $G/J(G)$ is the largest quotient that is a Cartesian product of finite simple groups. Moreover, the series $G &gt; J(G) &gt; J(J(G)) &gt; \dots$ always has trivial intersection, so it can be a useful characteristic series, especially if the terms are known to have finite index (equivalently, if $J(H)$ has finite index for every open subgroup $H$). Zalesskii showed that if $G$ is an infinite profinite group and $J(H)$ has finite index for every open subgroup $H$, then $G$ has a just infinite quotient.</p> <p>If $G$ is a compactly generated locally compact group with no infinite discrete quotient, then Caprace--Monod showed that there is a smallest cocompact normal subgroup $N$ of $G$, and $N/J(N)$ is generated topologically by a finite set of topologically simple closed normal subgroups, with $N &gt; J(N)$ unless $N$ is connected and soluble.</p>
Is the exponential function the sole solution to these equations?
260,298
13
https://mathoverflow.net/questions/260298/is-the-exponential-function-the-sole-solution-to-these-equations
<p>Let us take the exponential function $\lambda^z$ where $0 &lt; \lambda &lt; 1$. There are many great uniqueness conditions this holomorphic function satisfies. For example, it is the only function holomorphic in the right half plane, bounded in the right half plane, that interpolates $\lambda^{z} \Big{|}_{\mathbb{N}}$. It is also the only function to satisfy the differential equation $\frac{d}{dz}f(z) = \log(\lambda)f(z)$ when $f(0) =1$. Also it is the only function such that $f(z)f(w) = f(z+w)$ and $f(1) = \lambda$ and $f:\mathbb{R}^+\to\mathbb{R}$. It is also the only function holomorphic such that $f(z)f(w) = f(z+w)$ and $f(1) = \lambda$ and $f$ has a period of $2 \pi i /\log(\lambda)$.</p> <p>I'm asking this question because of this similar question on tetration <a href="https://mathoverflow.net/questions/259278/an-explicit-series-representation-for-the-analytic-tetration-with-complex-height">An explicit series representation for the analytic tetration with complex height</a> However when approaching the uniqueness of this tetration function on the real positive line, what is barring our path is a lemma (if you will) about the exponential function. This lemma is rather simple to state, and leads me to believe there must be a solution of this question.</p> <p>I am not conjecturing this lemma to be true by calling it a lemma (I have a feeling its true); it'd just be nice to see a proof in the positive. If someone has references or suggestions or even a straight proof about why this must be true/not true, that'd be outstanding. The lemma has to do with completely monotonic functions, and is rather tricky. It can be stated plainly as</p> <p>If a function $F$ is analytic on $\mathbb{R}^+$ and satisfies</p> <p>$$F(1) =\lambda$$ $$\lambda F(x) = F(x+1)$$ $$(-1)^n\frac{d^n}{dx^n}F(x) &gt;0$$</p> <p>must $F(x) = \lambda^x$?</p> <p>This can also be stated, if $F(x) = \lambda^{x}\phi(x)$ for some $1$-periodic function $\phi(x)$, and $(-1)^n\frac{d^n}{dx^n}F(x) &gt; 0$, must $\phi(x)$ be constant?</p> <p>It can also be stated in the manner I think is most likely to have an answer ready at hand. I also believe this to be the most valuable way of stating the question.</p> <p>Letting $F$ be analytic on $\mathbb{R}^+$, if $F(1) = e$ and $$\frac{d^n}{dx^n}F(x) &gt; 0$$ $$e\cdot F(x)= F(x+1)$$ must $F(x) = e^{x}$?</p>
[ { "answer_id": 260304, "score": 18, "text": "<p>The answer is yes. Let me address the last stated form of the question.\nFunctions $F$ with $F^{(n)}\\geq 0$ are called totally monotone, and S. Bernstein\nproved that all such functions are Laplace transforms of positive measures,\nthat is \n$$F(x)=\\int e^{xt}d\\mu(t).$$\nYour functional equation $F(x+1)=eF(x)$ then implies\n$$\\int e^{xt}(e^td\\mu(t))=\\int e^{xt}(ed\\mu(t)).$$\nFrom this follows $e^td\\mu(t)=ed\\mu(t)$ by the uniqueness theorem for the Laplace transform. This means that $\\mu$ is an atom at the point $1$. \nThis means that $F(x)=ce^x$, and your normalization implies $c=1$.</p>\n\n<p>The references for Laplace transform are Widder, Laplace transform, and Feller,\nAn introduction to probability... vol. II. Widder has Bernstein's theorem and\nuniqueness theorem.</p>\n" }, { "answer_id": 260307, "score": 9, "text": "<p>Here's a proof that also uses Bernstein's representation theorem for totally monotone functions (I wondered if that result was needed, and Alexander Eremenko's answer now makes me think, that it is). </p>\n\n<p>If $F:\\mathbb{R}_+\\to\\mathbb{R}_+$ satisfies $F(x+1)=eF(x)$, it is of the form $F(x)=\\exp(x+h(x))$, with $h$ a $1$-periodic function. But a totally monotonic function is logarithmically convex (see below) and $x+h(x)$ can only be convex if $h$ is a constant (actually zero, if also $F(0)=1$ or $F(1)=e$).</p>\n\n<p>$$*$$</p>\n\n<p><em>Totally monotonic functions are logarithmically convex:</em></p>\n\n<p>From $F(x)=\\int_{\\mathbb{R}_+}e^{\\lambda x}d\\mu(\\lambda)$, differentiating under the integral sign and applying the Cauchy-Schwarz inequality\n$$F'(x)^2=\\bigg(\\int_{\\mathbb{R}_+}\\lambda e^{\\lambda x}d\\mu(\\lambda)\\bigg)^2=\\bigg(\\int_{\\mathbb{R}_+}\\lambda e^{\\lambda x/2}\\cdot e^{\\lambda x/2}d\\mu(\\lambda)\\bigg)^2\\le$$$$\n\\int_{\\mathbb{R}_+}\\lambda^2 e^{\\lambda x}d\\mu(\\lambda)\\int_{\\mathbb{R}_+}e^{\\lambda x}d\\mu(\\lambda)=F''(x)F(x)\\ . $$</p>\n\n<p>Therefore $(F'/F)'\\ge0,$ that is $\\log F(x)$ is convex.</p>\n" }, { "answer_id": 260304, "score": 18, "text": "<p>The answer is yes. Let me address the last stated form of the question.\nFunctions $F$ with $F^{(n)}\\geq 0$ are called totally monotone, and S. Bernstein\nproved that all such functions are Laplace transforms of positive measures,\nthat is \n$$F(x)=\\int e^{xt}d\\mu(t).$$\nYour functional equation $F(x+1)=eF(x)$ then implies\n$$\\int e^{xt}(e^td\\mu(t))=\\int e^{xt}(ed\\mu(t)).$$\nFrom this follows $e^td\\mu(t)=ed\\mu(t)$ by the uniqueness theorem for the Laplace transform. This means that $\\mu$ is an atom at the point $1$. \nThis means that $F(x)=ce^x$, and your normalization implies $c=1$.</p>\n\n<p>The references for Laplace transform are Widder, Laplace transform, and Feller,\nAn introduction to probability... vol. II. Widder has Bernstein's theorem and\nuniqueness theorem.</p>\n" }, { "answer_id": 260307, "score": 9, "text": "<p>Here's a proof that also uses Bernstein's representation theorem for totally monotone functions (I wondered if that result was needed, and Alexander Eremenko's answer now makes me think, that it is). </p>\n\n<p>If $F:\\mathbb{R}_+\\to\\mathbb{R}_+$ satisfies $F(x+1)=eF(x)$, it is of the form $F(x)=\\exp(x+h(x))$, with $h$ a $1$-periodic function. But a totally monotonic function is logarithmically convex (see below) and $x+h(x)$ can only be convex if $h$ is a constant (actually zero, if also $F(0)=1$ or $F(1)=e$).</p>\n\n<p>$$*$$</p>\n\n<p><em>Totally monotonic functions are logarithmically convex:</em></p>\n\n<p>From $F(x)=\\int_{\\mathbb{R}_+}e^{\\lambda x}d\\mu(\\lambda)$, differentiating under the integral sign and applying the Cauchy-Schwarz inequality\n$$F'(x)^2=\\bigg(\\int_{\\mathbb{R}_+}\\lambda e^{\\lambda x}d\\mu(\\lambda)\\bigg)^2=\\bigg(\\int_{\\mathbb{R}_+}\\lambda e^{\\lambda x/2}\\cdot e^{\\lambda x/2}d\\mu(\\lambda)\\bigg)^2\\le$$$$\n\\int_{\\mathbb{R}_+}\\lambda^2 e^{\\lambda x}d\\mu(\\lambda)\\int_{\\mathbb{R}_+}e^{\\lambda x}d\\mu(\\lambda)=F''(x)F(x)\\ . $$</p>\n\n<p>Therefore $(F'/F)'\\ge0,$ that is $\\log F(x)$ is convex.</p>\n" }, { "answer_id": 260304, "score": 18, "text": "<p>The answer is yes. Let me address the last stated form of the question.\nFunctions $F$ with $F^{(n)}\\geq 0$ are called totally monotone, and S. Bernstein\nproved that all such functions are Laplace transforms of positive measures,\nthat is \n$$F(x)=\\int e^{xt}d\\mu(t).$$\nYour functional equation $F(x+1)=eF(x)$ then implies\n$$\\int e^{xt}(e^td\\mu(t))=\\int e^{xt}(ed\\mu(t)).$$\nFrom this follows $e^td\\mu(t)=ed\\mu(t)$ by the uniqueness theorem for the Laplace transform. This means that $\\mu$ is an atom at the point $1$. \nThis means that $F(x)=ce^x$, and your normalization implies $c=1$.</p>\n\n<p>The references for Laplace transform are Widder, Laplace transform, and Feller,\nAn introduction to probability... vol. II. Widder has Bernstein's theorem and\nuniqueness theorem.</p>\n" }, { "answer_id": 260307, "score": 9, "text": "<p>Here's a proof that also uses Bernstein's representation theorem for totally monotone functions (I wondered if that result was needed, and Alexander Eremenko's answer now makes me think, that it is). </p>\n\n<p>If $F:\\mathbb{R}_+\\to\\mathbb{R}_+$ satisfies $F(x+1)=eF(x)$, it is of the form $F(x)=\\exp(x+h(x))$, with $h$ a $1$-periodic function. But a totally monotonic function is logarithmically convex (see below) and $x+h(x)$ can only be convex if $h$ is a constant (actually zero, if also $F(0)=1$ or $F(1)=e$).</p>\n\n<p>$$*$$</p>\n\n<p><em>Totally monotonic functions are logarithmically convex:</em></p>\n\n<p>From $F(x)=\\int_{\\mathbb{R}_+}e^{\\lambda x}d\\mu(\\lambda)$, differentiating under the integral sign and applying the Cauchy-Schwarz inequality\n$$F'(x)^2=\\bigg(\\int_{\\mathbb{R}_+}\\lambda e^{\\lambda x}d\\mu(\\lambda)\\bigg)^2=\\bigg(\\int_{\\mathbb{R}_+}\\lambda e^{\\lambda x/2}\\cdot e^{\\lambda x/2}d\\mu(\\lambda)\\bigg)^2\\le$$$$\n\\int_{\\mathbb{R}_+}\\lambda^2 e^{\\lambda x}d\\mu(\\lambda)\\int_{\\mathbb{R}_+}e^{\\lambda x}d\\mu(\\lambda)=F''(x)F(x)\\ . $$</p>\n\n<p>Therefore $(F'/F)'\\ge0,$ that is $\\log F(x)$ is convex.</p>\n" } ]
Question: Is the exponential function the sole solution to these equations? Answer 1: <p>The answer is yes. Let me address the last stated form of the question. Functions $F$ with $F^{(n)}\geq 0$ are called totally monotone, and S. Bernstein proved that all such functions are Laplace transforms of positive measures, that is $$F(x)=\int e^{xt}d\mu(t).$$ Your functional equation $F(x+1)=eF(x)$ then implies $$\int e^{xt}(e^td\mu(t))=\int e^{xt}(ed\mu(t)).$$ From this follows $e^td\mu(t)=ed\mu(t)$ by the uniqueness theorem for the Laplace transform. This means that $\mu$ is an atom at the point $1$. This means that $F(x)=ce^x$, and your normalization implies $c=1$.</p> <p>The references for Laplace transform are Widder, Laplace transform, and Feller, An introduction to probability... vol. II. Widder has Bernstein's theorem and uniqueness theorem.</p> Answer 2: <p>Here's a proof that also uses Bernstein's representation theorem for totally monotone functions (I wondered if that result was needed, and Alexander Eremenko's answer now makes me think, that it is). </p> <p>If $F:\mathbb{R}_+\to\mathbb{R}_+$ satisfies $F(x+1)=eF(x)$, it is of the form $F(x)=\exp(x+h(x))$, with $h$ a $1$-periodic function. But a totally monotonic function is logarithmically convex (see below) and $x+h(x)$ can only be convex if $h$ is a constant (actually zero, if also $F(0)=1$ or $F(1)=e$).</p> <p>$$*$$</p> <p><em>Totally monotonic functions are logarithmically convex:</em></p> <p>From $F(x)=\int_{\mathbb{R}_+}e^{\lambda x}d\mu(\lambda)$, differentiating under the integral sign and applying the Cauchy-Schwarz inequality $$F'(x)^2=\bigg(\int_{\mathbb{R}_+}\lambda e^{\lambda x}d\mu(\lambda)\bigg)^2=\bigg(\int_{\mathbb{R}_+}\lambda e^{\lambda x/2}\cdot e^{\lambda x/2}d\mu(\lambda)\bigg)^2\le$$$$ \int_{\mathbb{R}_+}\lambda^2 e^{\lambda x}d\mu(\lambda)\int_{\mathbb{R}_+}e^{\lambda x}d\mu(\lambda)=F''(x)F(x)\ . $$</p> <p>Therefore $(F'/F)'\ge0,$ that is $\log F(x)$ is convex.</p> Answer 3: <p>The answer is yes. Let me address the last stated form of the question. Functions $F$ with $F^{(n)}\geq 0$ are called totally monotone, and S. Bernstein proved that all such functions are Laplace transforms of positive measures, that is $$F(x)=\int e^{xt}d\mu(t).$$ Your functional equation $F(x+1)=eF(x)$ then implies $$\int e^{xt}(e^td\mu(t))=\int e^{xt}(ed\mu(t)).$$ From this follows $e^td\mu(t)=ed\mu(t)$ by the uniqueness theorem for the Laplace transform. This means that $\mu$ is an atom at the point $1$. This means that $F(x)=ce^x$, and your normalization implies $c=1$.</p> <p>The references for Laplace transform are Widder, Laplace transform, and Feller, An introduction to probability... vol. II. Widder has Bernstein's theorem and uniqueness theorem.</p> Answer 4: <p>Here's a proof that also uses Bernstein's representation theorem for totally monotone functions (I wondered if that result was needed, and Alexander Eremenko's answer now makes me think, that it is). </p> <p>If $F:\mathbb{R}_+\to\mathbb{R}_+$ satisfies $F(x+1)=eF(x)$, it is of the form $F(x)=\exp(x+h(x))$, with $h$ a $1$-periodic function. But a totally monotonic function is logarithmically convex (see below) and $x+h(x)$ can only be convex if $h$ is a constant (actually zero, if also $F(0)=1$ or $F(1)=e$).</p> <p>$$*$$</p> <p><em>Totally monotonic functions are logarithmically convex:</em></p> <p>From $F(x)=\int_{\mathbb{R}_+}e^{\lambda x}d\mu(\lambda)$, differentiating under the integral sign and applying the Cauchy-Schwarz inequality $$F'(x)^2=\bigg(\int_{\mathbb{R}_+}\lambda e^{\lambda x}d\mu(\lambda)\bigg)^2=\bigg(\int_{\mathbb{R}_+}\lambda e^{\lambda x/2}\cdot e^{\lambda x/2}d\mu(\lambda)\bigg)^2\le$$$$ \int_{\mathbb{R}_+}\lambda^2 e^{\lambda x}d\mu(\lambda)\int_{\mathbb{R}_+}e^{\lambda x}d\mu(\lambda)=F''(x)F(x)\ . $$</p> <p>Therefore $(F'/F)'\ge0,$ that is $\log F(x)$ is convex.</p> Answer 5: <p>The answer is yes. Let me address the last stated form of the question. Functions $F$ with $F^{(n)}\geq 0$ are called totally monotone, and S. Bernstein proved that all such functions are Laplace transforms of positive measures, that is $$F(x)=\int e^{xt}d\mu(t).$$ Your functional equation $F(x+1)=eF(x)$ then implies $$\int e^{xt}(e^td\mu(t))=\int e^{xt}(ed\mu(t)).$$ From this follows $e^td\mu(t)=ed\mu(t)$ by the uniqueness theorem for the Laplace transform. This means that $\mu$ is an atom at the point $1$. This means that $F(x)=ce^x$, and your normalization implies $c=1$.</p> <p>The references for Laplace transform are Widder, Laplace transform, and Feller, An introduction to probability... vol. II. Widder has Bernstein's theorem and uniqueness theorem.</p> Answer 6: <p>Here's a proof that also uses Bernstein's representation theorem for totally monotone functions (I wondered if that result was needed, and Alexander Eremenko's answer now makes me think, that it is). </p> <p>If $F:\mathbb{R}_+\to\mathbb{R}_+$ satisfies $F(x+1)=eF(x)$, it is of the form $F(x)=\exp(x+h(x))$, with $h$ a $1$-periodic function. But a totally monotonic function is logarithmically convex (see below) and $x+h(x)$ can only be convex if $h$ is a constant (actually zero, if also $F(0)=1$ or $F(1)=e$).</p> <p>$$*$$</p> <p><em>Totally monotonic functions are logarithmically convex:</em></p> <p>From $F(x)=\int_{\mathbb{R}_+}e^{\lambda x}d\mu(\lambda)$, differentiating under the integral sign and applying the Cauchy-Schwarz inequality $$F'(x)^2=\bigg(\int_{\mathbb{R}_+}\lambda e^{\lambda x}d\mu(\lambda)\bigg)^2=\bigg(\int_{\mathbb{R}_+}\lambda e^{\lambda x/2}\cdot e^{\lambda x/2}d\mu(\lambda)\bigg)^2\le$$$$ \int_{\mathbb{R}_+}\lambda^2 e^{\lambda x}d\mu(\lambda)\int_{\mathbb{R}_+}e^{\lambda x}d\mu(\lambda)=F''(x)F(x)\ . $$</p> <p>Therefore $(F'/F)'\ge0,$ that is $\log F(x)$ is convex.</p>
Conjecture about harmonic numbers
260,516
12
https://mathoverflow.net/questions/260516/conjecture-about-harmonic-numbers
<p>I was inspired by <a href="https://math.stackexchange.com/questions/2062960/there-exist-infinite-many-n-in-mathbbn-such-that-s-n-s-n-frac1n2?noredirect=1#comment4336226_2062960">this</a> topic on Math.SE.<br> Suppose that $H_n = \sum\limits_{k=1}^n \frac{1}{k}$ - $n$th harmonic number. Then</p> <h2>Conjecture</h2> <blockquote> <p>Let $M$ be a set of all $n$ such that $$H_n - \lfloor{H_n\rfloor} &lt; \frac{1}{n^{1+\epsilon}}.$$ Then $$\forall\epsilon&gt;0 : |M| = \bar\eta(\epsilon) &lt; \infty.$$</p> </blockquote> <p>Picture below illustrates my conjecture, where I have checked this conjecture for $n &lt; 10^6$ for each $\epsilon\in(0,1.1)$ with step 0.01 <a href="https://i.stack.imgur.com/ArpSz.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/ArpSz.png" alt="enter image description here"></a> </p> <p>The following picture based on data provided by @GottfriedHelms for $n \approx 10^{100}$ (see answer below). <a href="https://i.stack.imgur.com/ySVSZ.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/ySVSZ.png" alt="enter image description here"></a></p>
[ { "answer_id": 260669, "score": 4, "text": "<p><strong><em>This is a comment at the comments of Gerhard \"Still Computing Oh So Slowly\" Paseman's answer, giving just indexes n for more record-holders.</em></strong> </p>\n\n<p>Let $\\small h_n$ denote the <em>n</em>'th harmonic number, $\\small A_n=\\{h_n\\}$ its fractional part. The sequence of $A_n$ has a remarkable shape like a sawtooth-curve with increasing wavelength when <em>n</em> is increased and with sharp local minima - a shape which can be exploited when we seek for possible <em>n</em> to be included in <strong><em>M</em></strong>.<br>\nI use $\\small f(n) = w_n = \\{h_n\\} \\cdot n$ and $\\small f(n,\\varepsilon)=w_n \\cdot n^\\varepsilon $ with the OP's condition rewritten as $\\small f(n,\\varepsilon)&lt;1 $ as criterion for the inclusion of <em>n</em> into the set <strong><em>M</em></strong> . Of course the cardinality of <strong><em>M</em></strong> is limited by an upper bound $ \\small n \\le N$ with some <strong><em>N</em></strong> that can numerically be handled, so actually we should write explicitely $\\small |M(\\varepsilon,N)| $ instead of <strong><em>M</em></strong> only. I could manage to use $\\small N \\approx e^{2000}$ with the help of Pari/GP. </p>\n\n<p>The local minima of $\\small A_n$ occur near $\\small x_k=e^{k-\\gamma}, k \\in \\mathbb N^+$ and the <em>n</em> to be tested is one of the next integers enclosing the $\\small x_k$ - so this exact <em>n</em> must be empirically be determined. (Note that my <em>n</em> are the <em>n+1</em> used in the comments at the OP denoting the previous <em>high</em> value of <strong><em>A</em></strong> ) </p>\n\n<p>The harmonic numbers $\\small h_n$ can be computed in Pari/GP using <code>h(n) = psi(1+n) + Euler</code> ; however, this seems to be limited to something like $ \\small n \\lt e^{600}$ and so I had to introduce the Euler-McLaurin-formula for the larger <em>n</em> and implemented the switch from one method to the other at <code>n = 1e50</code> </p>\n\n<p><hr>\nThe following table shows the first <em>40</em> entries of my <em>2000</em>-row table containing $\\small n \\approx 10^{840}$ which was possible to compute using Pari/GP in just a couple of seconds. </p>\n\n<p>The basic option according to the focus in the OP's question is to test $\\small f(n,\\varepsilon) = w_n \\cdot n^\\varepsilon \\lt 1$. For an example with $\\small \\varepsilon=0.1$ see the 6'th column and the 7'th column allowing to sum to the cardinality of <strong><em>M</em></strong> where we find $\\small | M(0.1,2000) | = 7$ with $\\small n \\le N \\approx e^{2000}$ </p>\n\n<p>A second, much nicer, option is to define the function $\\small e(n) = -\\log_n(w_n) $ and test $\\small e(n) \\gt \\varepsilon$ whether to include this <em>n</em> or not. This is simply possible when looking into the <em>5</em>'th column and simply compare.<br>\nThis latter method allows to compute the cardinalities of $\\small M(\\varepsilon)$ for arbitrary $\\varepsilon$ really fast, for instance to create informative scatter- or lineplots, when first a list for $e(n)$ of length $\\small N$ is created and then the successful comparisions with the intended $\\varepsilon$ are summed to determine the cardinality. </p>\n\n<pre><code> n | h_n | A=frac(h_n) | w=A*n | e(n) | w*n^0.1|in M?\n -----------------+--------+-------------+----------+-----------+--------+---\n 1 1.00000 0.00000 0.00000 1.000000 0.00000 1\n 4 2.08333 0.0833333 0.333333 0.792481 0.382899 1\n 11 3.01988 0.0198773 0.218651 0.634006 0.277901 1\n 31 4.02725 0.0272452 0.844601 0.0491822 1.19066 .\n 83 5.00207 0.00206827 0.171667 0.398793 0.267051 1\n 227 6.00437 0.00436671 0.991243 0.00162136 1.70523 .\n 616 7.00127 0.00127410 0.784844 0.0377178 1.49191 .\n 1674 8.00049 0.000485572 0.812848 0.0279149 1.70759 .\n 4550 9.00021 0.000208063 0.946686 0.00650460 2.19790 .\n 12367 10.0000 0.0000430083 0.531883 0.0670005 1.36472 .\n 33617 11.0000 0.0000177086 0.595311 0.0497632 1.68811 .\n 91380 12.0000 0.00000305167 0.278861 0.111798 0.873923 1\n 248397 13.0000 0.00000122948 0.305399 0.0954806 1.05775 .\n 675214 14.0000 1.36205E-7 0.919678 0.00623806 3.52030 .\n 1835421 15.0000 3.78268E-7 0.694281 0.0252988 2.93703 .\n 4989191 16.0000 9.54538E-7 0.476237 0.0481002 2.22652 .\n 13562027 17.0000 1.48499E-8 0.201395 0.0975770 1.04059 .\n 36865412 18.0000 3.71993E-9 0.137137 0.114033 0.783098 1\n 100210581 19.0000 9.73330E-9 0.975380 0.00135314 6.15552 .\n 272400600 20.0000 1.61744E-9 0.440592 0.0421997 3.07297 .\n 740461601 21.0000 4.01333E-9 0.297172 0.0594162 2.29065 .\n 2012783315 22.0000 1.38447E-10 0.278664 0.0596444 2.37389 .\n 5471312310 23.0000 1.97920E-11 0.108288 0.0991384 1.01951 .\n 14872568831 24.0000 2.27220E-11 0.337935 0.0463183 3.51618 .\n 40427833596 25.0000 6.07937E-12 0.245776 0.0574601 2.82623 .\n 109894245429 26.0000 7.60776E-12 0.836049 0.00704359 10.6250 .\n 298723530401 27.0000 1.82203E-12 0.544283 0.0230213 7.64454 .\n 812014744422 28.0000 5.52830E-13 0.448906 0.0292071 6.96806 .\n 2207284924203 29.0000 1.00870E-13 0.222650 0.0528504 3.81951 .\n 6000022499693 30.0000 2.16954E-14 0.130173 0.0692963 2.46795 .\n 16309752131262 31.0000 3.65111E-14 0.595487 0.0170391 12.4772 .\n 44334502845080 32.0000 1.81005E-15 0.080247 0.0802804 1.85827 .\n 120513673457548 33.0000 4.59281E-15 0.553496 0.0182434 14.1651 .\n 327590128640500 34.0000 2.31992E-15 0.759983 0.00821173 21.4950 .\n 890482293866031 35.0000 2.71425E-17 0.024169 0.108145 0.755506 1\n 2420581837980561 36.0000 2.55560E-16 0.618603 0.0135588 21.3700 .\n 6579823624480555 37.0000 1.20561E-17 0.079326 0.0695767 3.02860 .\n 17885814992891026 38.0000 4.26642E-18 0.076308 0.0687541 3.21976 .\n 48618685882356024 39.0000 7.23781E-19 0.035189 0.0871101 1.64093 .\n132159290357566703 40.0000 2.02186E-18 0.267208 0.0334763 13.7708 .\n</code></pre>\n\n<p>The set $\\small M(0.1,e^{40})$ (represented by <em>40</em> rows) is here the set of all $n$ where $\\small f(n,0.1)=w_n \\cdot n^{0.1} \\lt 1$ . The number of such entries for $\\small n=132159290357566703 \\approx 10^{17}$ is here <em>7</em> . </p>\n\n<p>All numerical tests which I've done indicate that the cardinality of $\\small M(\\varepsilon \\gt 0,\\infty)$ is finite and roughly reciprocal to $\\small \\varepsilon$ and only if $\\varepsilon=0$ is surely infinite. </p>\n\n<p><hr></p>\n\n<h1>Pari/GP tools</h1>\n\n<p>This is the Pari/GP-program which I used. </p>\n\n<p>Using the functions</p>\n\n<pre><code> h(n) = Euler+if(n&lt;1e50, psi(1+n),log(n)+1/2/n-1/12/n^2+1/120/n^4-1/252/n^6+1/240/n^8)\n A(n) = if(n==1,return(0)); frac(h(n))\n w(n) = A(n)*n \n e(n) = if(n==1,return(1)); -log(w(n))/log(n)\n</code></pre>\n\n<p>The following needs only two steps to find the <em>n</em> with the next local minimum when called with index $\\small k \\in \\mathbb N$ : </p>\n\n<pre><code> {find_n(k)=local(n1,n2,a1,a2);\n n1=floor(exp(k-Euler)); n2=n1+1; a1 = A(n1);a2 = A(n2);\n if(a1&lt;a2,return(n1),return(n2)); \n } \n\n\n\n \\\\ create that list one time, then evaluate cardinality for various eps\n \\\\ using that same list\n {makeList(listlen=40)= local(list,n,w1,logn);\n list=matrix(listlen,4);\n list[1,]=[Euler,0,0,0];\n for(k=2,listlen,\n n=find_n(k); logn=log(n);\n list[k,]=[logn+Euler,n,w1=w(n),-log(w1)/logn];\n );\n return(list); }\n\n \\\\ compute cardinality with some eps\n cardM(eps,list) = sum(k=1,#list[,1],list[k,4]&gt;eps) \n\n \\\\ apply, note: for long lists we need high internal precision\n list = makeList(40) \\\\ the max N is here about e^40\n print(cardM(0.1,list) )\n</code></pre>\n\n<hr>\n\n<p><strong><em>[table 2]</em></strong>: This are sample data for the OP's plot of the cardinality of $\\small M(\\varepsilon)$ by the argument $\\varepsilon$ where $\\small N \\approx \\exp(250) \\approx 10^{108}$ : </p>\n\n<pre><code> eps | c=#M | r=c*eps | c= card(M) for n&lt;= N approx 10^108\n ------+------------------\n 0.01 104 1.040\n 0.02 55 1.100\n 0.03 36 1.080\n 0.04 29 1.160\n 0.05 21 1.050\n 0.06 16 0.960\n 0.07 12 0.840\n 0.08 12 0.960\n 0.09 10 0.900\n 0.10 7 0.700\n 0.11 6 0.660\n 0.12 4 0.480\n 0.13 4 0.520\n 0.14 4 0.560\n 0.15 4 0.600\n 0.16 4 0.640\n 0.17 4 0.680\n 0.18 4 0.720\n 0.19 4 0.760\n 0.20 4 0.800\n</code></pre>\n\n<p><hr></p>\n\n<h1>Pictures</h1>\n\n<p><strong><em>[Picture 1]</em></strong> : I've extended the search-space for $n$ to the range $\\small 1 \\ldots e^{1000} \\approx 10^{434} $ and to check sanity to the range $\\small 1 \\ldots e^{2000} \\approx 10^{868} $. For epsilons $\\varepsilon$ from $0.001$ in $200$ steps up to $0.2$ I made the following graph:<br>\n<a href=\"https://i.stack.imgur.com/DAz5g.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/DAz5g.png\" alt=\"picture\"></a> </p>\n\n<p>Remarks: for the very small epsilon the increase of the search-space gives slightly higher results, which also illustrates, that for \"larger\" epsilon the cardinality of $\\small M(\\varepsilon)$ is finite</p>\n\n<p><hr>\n<strong><em>[Picture 2]</em></strong>: Indicates uniformity of the $\\small f(n) =w_n = frac(h_n) \\cdot n^1$ at the <em>n</em>, where $\\small w_n $ has a local minimum (on request of @GerhardPaseman):<br>\n<a href=\"https://i.stack.imgur.com/18Az3.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/18Az3.png\" alt=\"picture 2\"></a> </p>\n\n<p><strong><em>[Picture 3]</em></strong>:It is also convenient to show a rescaling of the $\\small f(n)$ so that we can immediately determine the cardinalities $\\small |M(\\varepsilon)|$ just by counting the number of dots $\\small e(n)$ above $\\small \\varepsilon$. The derivation of the formula is </p>\n\n<p>$$ \\small{ \\begin{array}{lll}\n A(n) &amp;\\lt &amp; {1\\over n^1\\cdot n^\\varepsilon} &amp; \\text{ from OP}\\\\\n f(n) \\cdot n^\\varepsilon &amp;\\lt &amp; 1\\\\\n \\ln(f(n)) + \\varepsilon \\cdot \\ln(n) &amp;\\lt &amp; 0 \\\\\n {\\ln(f(n)) \\over \\ln(n)} + \\varepsilon &amp;\\lt &amp; 0 \\\\\n \\varepsilon &amp; \\lt&amp; -\\ln_n(f(n)) \n\\end{array} }$$</p>\n\n<p>and we simply count, how many dots in the picture show $\\small e(n) = - \\ln_n(f(n)) \\gt \\varepsilon$ </p>\n\n<p><a href=\"https://i.stack.imgur.com/QFKcS.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/QFKcS.png\" alt=\"picture\"></a> </p>\n\n<p><hr></p>\n\n<h1>Data for cardinalities-plot</h1>\n\n<pre><code>epsilon |M(eps,N)||M(eps,N)| \n N~e^1000 N~e^2000\n-------------------------------\n0.0000 1000 2000\n0.0010 633 862\n0.0020 430 482\n0.0030 315 330\n0.0040 258 264\n0.0050 204 205\n0.0060 171 171\n0.0070 151 151\n0.0080 135 135\n0.0090 120 120\n0.0100 109 109\n0.0110 100 100\n0.0120 98 98\n0.0130 89 89\n0.0140 84 84\n0.0150 79 79\n0.0160 73 73\n0.0170 69 69\n0.0180 62 62\n0.0190 57 57\n0.0200 55 55\n0.0210 50 50\n0.0220 47 47\n0.0230 44 44\n0.0240 42 42\n0.0250 41 41\n0.0260 39 39\n0.0270 39 39\n0.0280 37 37\n0.0290 37 37\n0.0300 36 36\n0.0310 36 36\n0.0320 36 36\n0.0330 36 36\n0.0340 34 34\n0.0350 34 34\n0.0360 34 34\n0.0370 33 33\n0.0380 32 32\n0.0390 31 31\n0.0400 29 29\n0.0410 29 29\n0.0420 29 29\n0.0430 26 26\n0.0440 26 26\n0.0450 25 25\n0.0460 25 25\n0.0470 24 24\n0.0480 24 24\n0.0490 23 23\n0.0500 21 21\n0.0510 21 21\n0.0520 21 21\n0.0530 20 20\n0.0540 20 20\n0.0550 20 20\n0.0560 20 20\n0.0570 19 19\n0.0580 18 18\n0.0590 18 18\n0.0600 16 16\n0.0610 16 16\n0.0620 16 16\n0.0630 16 16\n0.0640 16 16\n0.0650 16 16\n0.0660 16 16\n0.0670 16 16\n0.0680 15 15\n0.0690 14 14\n0.0700 12 12\n0.0710 12 12\n0.0720 12 12\n0.0730 12 12\n0.0740 12 12\n0.0750 12 12\n0.0760 12 12\n0.0770 12 12\n0.0780 12 12\n0.0790 12 12\n0.0800 12 12\n0.0810 11 11\n0.0820 11 11\n0.0830 11 11\n0.0840 11 11\n0.0850 11 11\n0.0860 11 11\n0.0870 11 11\n0.0880 10 10\n0.0890 10 10\n0.0900 10 10\n0.0910 10 10\n0.0920 10 10\n0.0930 10 10\n0.0940 10 10\n0.0950 10 10\n0.0960 9 9\n0.0970 9 9\n0.0980 8 8\n0.0990 8 8\n0.1000 7 7\n0.1010 7 7\n0.1020 7 7\n0.1030 7 7\n0.1040 7 7\n0.1050 7 7\n0.1060 7 7\n0.1070 7 7\n0.1080 7 7\n0.1090 6 6\n0.1100 6 6\n0.1110 6 6\n0.1120 5 5\n0.1130 5 5\n0.1140 5 5\n0.1150 4 4\n0.1160 4 4\n0.1170 4 4\n0.1180 4 4\n0.1190 4 4\n0.1200 4 4\n0.1210 4 4\n0.1220 4 4\n0.1230 4 4\n0.1240 4 4\n0.1250 4 4\n0.1260 4 4\n0.1270 4 4\n0.1280 4 4\n0.1290 4 4\n0.1300 4 4\n0.1310 4 4\n0.1320 4 4\n0.1330 4 4\n0.1340 4 4\n0.1350 4 4\n0.1360 4 4\n0.1370 4 4\n0.1380 4 4\n0.1390 4 4\n0.1400 4 4\n0.1410 4 4\n0.1420 4 4\n0.1430 4 4\n0.1440 4 4\n0.1450 4 4\n0.1460 4 4\n0.1470 4 4\n0.1480 4 4\n0.1490 4 4\n0.1500 4 4\n0.1510 4 4\n0.1520 4 4\n0.1530 4 4\n0.1540 4 4\n0.1550 4 4\n0.1560 4 4\n0.1570 4 4\n0.1580 4 4\n0.1590 4 4\n0.1600 4 4\n0.1610 4 4\n0.1620 4 4\n0.1630 4 4\n0.1640 4 4\n0.1650 4 4\n0.1660 4 4\n0.1670 4 4\n0.1680 4 4\n0.1690 4 4\n0.1700 4 4\n0.1710 4 4\n0.1720 4 4\n0.1730 4 4\n0.1740 4 4\n0.1750 4 4\n0.1760 4 4\n0.1770 4 4\n0.1780 4 4\n0.1790 4 4\n0.1800 4 4\n0.1810 4 4\n0.1820 4 4\n0.1830 4 4\n0.1840 4 4\n0.1850 4 4\n0.1860 4 4\n0.1870 4 4\n0.1880 4 4\n0.1890 4 4\n0.1900 4 4\n0.1910 4 4\n0.1920 4 4\n0.1930 4 4\n0.1940 4 4\n0.1950 4 4\n0.1960 4 4\n0.1970 4 4\n0.1980 4 4\n0.1990 4 4\n</code></pre>\n" }, { "answer_id": 260571, "score": 2, "text": "<p>I am letting $A = H_n - \\lfloor H_n \\rfloor$ and computing it in a dumb way, and then computing $1/nA$ when $A \\lt 1/n $. For $n=83$ I get $1/nA \\approx 5.825$, which is the largest such value I find so far, and which means 83 is one of three values of $n$ which are allowed when $\\epsilon$ is 0.3. A smarter way to compute the values of $n$ for which $1/nA$ is worth computing is to use $\\log(n+1/2) + \\gamma$ for just those $n$ for which the expression is very near an integer, for $H_n$ is within $1/24n^2$ of this value. $\\log(1/nA )/\\log(n)$ then serves as an upper bound on $\\epsilon$.</p>\n\n<p>My calculations show $1 \\lt 1/nA \\lt 2$ much of the time, which means $\\epsilon \\lt 1/\\log n$ for admissible $n$. I expect the conjecture to hold and when proved to reveal some of the nature of Euler's $\\gamma$.</p>\n\n<p>Gerhard \"Still Computing Oh So Slowly\" Paseman, 2017.01.25.</p>\n" }, { "answer_id": 260669, "score": 4, "text": "<p><strong><em>This is a comment at the comments of Gerhard \"Still Computing Oh So Slowly\" Paseman's answer, giving just indexes n for more record-holders.</em></strong> </p>\n\n<p>Let $\\small h_n$ denote the <em>n</em>'th harmonic number, $\\small A_n=\\{h_n\\}$ its fractional part. The sequence of $A_n$ has a remarkable shape like a sawtooth-curve with increasing wavelength when <em>n</em> is increased and with sharp local minima - a shape which can be exploited when we seek for possible <em>n</em> to be included in <strong><em>M</em></strong>.<br>\nI use $\\small f(n) = w_n = \\{h_n\\} \\cdot n$ and $\\small f(n,\\varepsilon)=w_n \\cdot n^\\varepsilon $ with the OP's condition rewritten as $\\small f(n,\\varepsilon)&lt;1 $ as criterion for the inclusion of <em>n</em> into the set <strong><em>M</em></strong> . Of course the cardinality of <strong><em>M</em></strong> is limited by an upper bound $ \\small n \\le N$ with some <strong><em>N</em></strong> that can numerically be handled, so actually we should write explicitely $\\small |M(\\varepsilon,N)| $ instead of <strong><em>M</em></strong> only. I could manage to use $\\small N \\approx e^{2000}$ with the help of Pari/GP. </p>\n\n<p>The local minima of $\\small A_n$ occur near $\\small x_k=e^{k-\\gamma}, k \\in \\mathbb N^+$ and the <em>n</em> to be tested is one of the next integers enclosing the $\\small x_k$ - so this exact <em>n</em> must be empirically be determined. (Note that my <em>n</em> are the <em>n+1</em> used in the comments at the OP denoting the previous <em>high</em> value of <strong><em>A</em></strong> ) </p>\n\n<p>The harmonic numbers $\\small h_n$ can be computed in Pari/GP using <code>h(n) = psi(1+n) + Euler</code> ; however, this seems to be limited to something like $ \\small n \\lt e^{600}$ and so I had to introduce the Euler-McLaurin-formula for the larger <em>n</em> and implemented the switch from one method to the other at <code>n = 1e50</code> </p>\n\n<p><hr>\nThe following table shows the first <em>40</em> entries of my <em>2000</em>-row table containing $\\small n \\approx 10^{840}$ which was possible to compute using Pari/GP in just a couple of seconds. </p>\n\n<p>The basic option according to the focus in the OP's question is to test $\\small f(n,\\varepsilon) = w_n \\cdot n^\\varepsilon \\lt 1$. For an example with $\\small \\varepsilon=0.1$ see the 6'th column and the 7'th column allowing to sum to the cardinality of <strong><em>M</em></strong> where we find $\\small | M(0.1,2000) | = 7$ with $\\small n \\le N \\approx e^{2000}$ </p>\n\n<p>A second, much nicer, option is to define the function $\\small e(n) = -\\log_n(w_n) $ and test $\\small e(n) \\gt \\varepsilon$ whether to include this <em>n</em> or not. This is simply possible when looking into the <em>5</em>'th column and simply compare.<br>\nThis latter method allows to compute the cardinalities of $\\small M(\\varepsilon)$ for arbitrary $\\varepsilon$ really fast, for instance to create informative scatter- or lineplots, when first a list for $e(n)$ of length $\\small N$ is created and then the successful comparisions with the intended $\\varepsilon$ are summed to determine the cardinality. </p>\n\n<pre><code> n | h_n | A=frac(h_n) | w=A*n | e(n) | w*n^0.1|in M?\n -----------------+--------+-------------+----------+-----------+--------+---\n 1 1.00000 0.00000 0.00000 1.000000 0.00000 1\n 4 2.08333 0.0833333 0.333333 0.792481 0.382899 1\n 11 3.01988 0.0198773 0.218651 0.634006 0.277901 1\n 31 4.02725 0.0272452 0.844601 0.0491822 1.19066 .\n 83 5.00207 0.00206827 0.171667 0.398793 0.267051 1\n 227 6.00437 0.00436671 0.991243 0.00162136 1.70523 .\n 616 7.00127 0.00127410 0.784844 0.0377178 1.49191 .\n 1674 8.00049 0.000485572 0.812848 0.0279149 1.70759 .\n 4550 9.00021 0.000208063 0.946686 0.00650460 2.19790 .\n 12367 10.0000 0.0000430083 0.531883 0.0670005 1.36472 .\n 33617 11.0000 0.0000177086 0.595311 0.0497632 1.68811 .\n 91380 12.0000 0.00000305167 0.278861 0.111798 0.873923 1\n 248397 13.0000 0.00000122948 0.305399 0.0954806 1.05775 .\n 675214 14.0000 1.36205E-7 0.919678 0.00623806 3.52030 .\n 1835421 15.0000 3.78268E-7 0.694281 0.0252988 2.93703 .\n 4989191 16.0000 9.54538E-7 0.476237 0.0481002 2.22652 .\n 13562027 17.0000 1.48499E-8 0.201395 0.0975770 1.04059 .\n 36865412 18.0000 3.71993E-9 0.137137 0.114033 0.783098 1\n 100210581 19.0000 9.73330E-9 0.975380 0.00135314 6.15552 .\n 272400600 20.0000 1.61744E-9 0.440592 0.0421997 3.07297 .\n 740461601 21.0000 4.01333E-9 0.297172 0.0594162 2.29065 .\n 2012783315 22.0000 1.38447E-10 0.278664 0.0596444 2.37389 .\n 5471312310 23.0000 1.97920E-11 0.108288 0.0991384 1.01951 .\n 14872568831 24.0000 2.27220E-11 0.337935 0.0463183 3.51618 .\n 40427833596 25.0000 6.07937E-12 0.245776 0.0574601 2.82623 .\n 109894245429 26.0000 7.60776E-12 0.836049 0.00704359 10.6250 .\n 298723530401 27.0000 1.82203E-12 0.544283 0.0230213 7.64454 .\n 812014744422 28.0000 5.52830E-13 0.448906 0.0292071 6.96806 .\n 2207284924203 29.0000 1.00870E-13 0.222650 0.0528504 3.81951 .\n 6000022499693 30.0000 2.16954E-14 0.130173 0.0692963 2.46795 .\n 16309752131262 31.0000 3.65111E-14 0.595487 0.0170391 12.4772 .\n 44334502845080 32.0000 1.81005E-15 0.080247 0.0802804 1.85827 .\n 120513673457548 33.0000 4.59281E-15 0.553496 0.0182434 14.1651 .\n 327590128640500 34.0000 2.31992E-15 0.759983 0.00821173 21.4950 .\n 890482293866031 35.0000 2.71425E-17 0.024169 0.108145 0.755506 1\n 2420581837980561 36.0000 2.55560E-16 0.618603 0.0135588 21.3700 .\n 6579823624480555 37.0000 1.20561E-17 0.079326 0.0695767 3.02860 .\n 17885814992891026 38.0000 4.26642E-18 0.076308 0.0687541 3.21976 .\n 48618685882356024 39.0000 7.23781E-19 0.035189 0.0871101 1.64093 .\n132159290357566703 40.0000 2.02186E-18 0.267208 0.0334763 13.7708 .\n</code></pre>\n\n<p>The set $\\small M(0.1,e^{40})$ (represented by <em>40</em> rows) is here the set of all $n$ where $\\small f(n,0.1)=w_n \\cdot n^{0.1} \\lt 1$ . The number of such entries for $\\small n=132159290357566703 \\approx 10^{17}$ is here <em>7</em> . </p>\n\n<p>All numerical tests which I've done indicate that the cardinality of $\\small M(\\varepsilon \\gt 0,\\infty)$ is finite and roughly reciprocal to $\\small \\varepsilon$ and only if $\\varepsilon=0$ is surely infinite. </p>\n\n<p><hr></p>\n\n<h1>Pari/GP tools</h1>\n\n<p>This is the Pari/GP-program which I used. </p>\n\n<p>Using the functions</p>\n\n<pre><code> h(n) = Euler+if(n&lt;1e50, psi(1+n),log(n)+1/2/n-1/12/n^2+1/120/n^4-1/252/n^6+1/240/n^8)\n A(n) = if(n==1,return(0)); frac(h(n))\n w(n) = A(n)*n \n e(n) = if(n==1,return(1)); -log(w(n))/log(n)\n</code></pre>\n\n<p>The following needs only two steps to find the <em>n</em> with the next local minimum when called with index $\\small k \\in \\mathbb N$ : </p>\n\n<pre><code> {find_n(k)=local(n1,n2,a1,a2);\n n1=floor(exp(k-Euler)); n2=n1+1; a1 = A(n1);a2 = A(n2);\n if(a1&lt;a2,return(n1),return(n2)); \n } \n\n\n\n \\\\ create that list one time, then evaluate cardinality for various eps\n \\\\ using that same list\n {makeList(listlen=40)= local(list,n,w1,logn);\n list=matrix(listlen,4);\n list[1,]=[Euler,0,0,0];\n for(k=2,listlen,\n n=find_n(k); logn=log(n);\n list[k,]=[logn+Euler,n,w1=w(n),-log(w1)/logn];\n );\n return(list); }\n\n \\\\ compute cardinality with some eps\n cardM(eps,list) = sum(k=1,#list[,1],list[k,4]&gt;eps) \n\n \\\\ apply, note: for long lists we need high internal precision\n list = makeList(40) \\\\ the max N is here about e^40\n print(cardM(0.1,list) )\n</code></pre>\n\n<hr>\n\n<p><strong><em>[table 2]</em></strong>: This are sample data for the OP's plot of the cardinality of $\\small M(\\varepsilon)$ by the argument $\\varepsilon$ where $\\small N \\approx \\exp(250) \\approx 10^{108}$ : </p>\n\n<pre><code> eps | c=#M | r=c*eps | c= card(M) for n&lt;= N approx 10^108\n ------+------------------\n 0.01 104 1.040\n 0.02 55 1.100\n 0.03 36 1.080\n 0.04 29 1.160\n 0.05 21 1.050\n 0.06 16 0.960\n 0.07 12 0.840\n 0.08 12 0.960\n 0.09 10 0.900\n 0.10 7 0.700\n 0.11 6 0.660\n 0.12 4 0.480\n 0.13 4 0.520\n 0.14 4 0.560\n 0.15 4 0.600\n 0.16 4 0.640\n 0.17 4 0.680\n 0.18 4 0.720\n 0.19 4 0.760\n 0.20 4 0.800\n</code></pre>\n\n<p><hr></p>\n\n<h1>Pictures</h1>\n\n<p><strong><em>[Picture 1]</em></strong> : I've extended the search-space for $n$ to the range $\\small 1 \\ldots e^{1000} \\approx 10^{434} $ and to check sanity to the range $\\small 1 \\ldots e^{2000} \\approx 10^{868} $. For epsilons $\\varepsilon$ from $0.001$ in $200$ steps up to $0.2$ I made the following graph:<br>\n<a href=\"https://i.stack.imgur.com/DAz5g.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/DAz5g.png\" alt=\"picture\"></a> </p>\n\n<p>Remarks: for the very small epsilon the increase of the search-space gives slightly higher results, which also illustrates, that for \"larger\" epsilon the cardinality of $\\small M(\\varepsilon)$ is finite</p>\n\n<p><hr>\n<strong><em>[Picture 2]</em></strong>: Indicates uniformity of the $\\small f(n) =w_n = frac(h_n) \\cdot n^1$ at the <em>n</em>, where $\\small w_n $ has a local minimum (on request of @GerhardPaseman):<br>\n<a href=\"https://i.stack.imgur.com/18Az3.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/18Az3.png\" alt=\"picture 2\"></a> </p>\n\n<p><strong><em>[Picture 3]</em></strong>:It is also convenient to show a rescaling of the $\\small f(n)$ so that we can immediately determine the cardinalities $\\small |M(\\varepsilon)|$ just by counting the number of dots $\\small e(n)$ above $\\small \\varepsilon$. The derivation of the formula is </p>\n\n<p>$$ \\small{ \\begin{array}{lll}\n A(n) &amp;\\lt &amp; {1\\over n^1\\cdot n^\\varepsilon} &amp; \\text{ from OP}\\\\\n f(n) \\cdot n^\\varepsilon &amp;\\lt &amp; 1\\\\\n \\ln(f(n)) + \\varepsilon \\cdot \\ln(n) &amp;\\lt &amp; 0 \\\\\n {\\ln(f(n)) \\over \\ln(n)} + \\varepsilon &amp;\\lt &amp; 0 \\\\\n \\varepsilon &amp; \\lt&amp; -\\ln_n(f(n)) \n\\end{array} }$$</p>\n\n<p>and we simply count, how many dots in the picture show $\\small e(n) = - \\ln_n(f(n)) \\gt \\varepsilon$ </p>\n\n<p><a href=\"https://i.stack.imgur.com/QFKcS.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/QFKcS.png\" alt=\"picture\"></a> </p>\n\n<p><hr></p>\n\n<h1>Data for cardinalities-plot</h1>\n\n<pre><code>epsilon |M(eps,N)||M(eps,N)| \n N~e^1000 N~e^2000\n-------------------------------\n0.0000 1000 2000\n0.0010 633 862\n0.0020 430 482\n0.0030 315 330\n0.0040 258 264\n0.0050 204 205\n0.0060 171 171\n0.0070 151 151\n0.0080 135 135\n0.0090 120 120\n0.0100 109 109\n0.0110 100 100\n0.0120 98 98\n0.0130 89 89\n0.0140 84 84\n0.0150 79 79\n0.0160 73 73\n0.0170 69 69\n0.0180 62 62\n0.0190 57 57\n0.0200 55 55\n0.0210 50 50\n0.0220 47 47\n0.0230 44 44\n0.0240 42 42\n0.0250 41 41\n0.0260 39 39\n0.0270 39 39\n0.0280 37 37\n0.0290 37 37\n0.0300 36 36\n0.0310 36 36\n0.0320 36 36\n0.0330 36 36\n0.0340 34 34\n0.0350 34 34\n0.0360 34 34\n0.0370 33 33\n0.0380 32 32\n0.0390 31 31\n0.0400 29 29\n0.0410 29 29\n0.0420 29 29\n0.0430 26 26\n0.0440 26 26\n0.0450 25 25\n0.0460 25 25\n0.0470 24 24\n0.0480 24 24\n0.0490 23 23\n0.0500 21 21\n0.0510 21 21\n0.0520 21 21\n0.0530 20 20\n0.0540 20 20\n0.0550 20 20\n0.0560 20 20\n0.0570 19 19\n0.0580 18 18\n0.0590 18 18\n0.0600 16 16\n0.0610 16 16\n0.0620 16 16\n0.0630 16 16\n0.0640 16 16\n0.0650 16 16\n0.0660 16 16\n0.0670 16 16\n0.0680 15 15\n0.0690 14 14\n0.0700 12 12\n0.0710 12 12\n0.0720 12 12\n0.0730 12 12\n0.0740 12 12\n0.0750 12 12\n0.0760 12 12\n0.0770 12 12\n0.0780 12 12\n0.0790 12 12\n0.0800 12 12\n0.0810 11 11\n0.0820 11 11\n0.0830 11 11\n0.0840 11 11\n0.0850 11 11\n0.0860 11 11\n0.0870 11 11\n0.0880 10 10\n0.0890 10 10\n0.0900 10 10\n0.0910 10 10\n0.0920 10 10\n0.0930 10 10\n0.0940 10 10\n0.0950 10 10\n0.0960 9 9\n0.0970 9 9\n0.0980 8 8\n0.0990 8 8\n0.1000 7 7\n0.1010 7 7\n0.1020 7 7\n0.1030 7 7\n0.1040 7 7\n0.1050 7 7\n0.1060 7 7\n0.1070 7 7\n0.1080 7 7\n0.1090 6 6\n0.1100 6 6\n0.1110 6 6\n0.1120 5 5\n0.1130 5 5\n0.1140 5 5\n0.1150 4 4\n0.1160 4 4\n0.1170 4 4\n0.1180 4 4\n0.1190 4 4\n0.1200 4 4\n0.1210 4 4\n0.1220 4 4\n0.1230 4 4\n0.1240 4 4\n0.1250 4 4\n0.1260 4 4\n0.1270 4 4\n0.1280 4 4\n0.1290 4 4\n0.1300 4 4\n0.1310 4 4\n0.1320 4 4\n0.1330 4 4\n0.1340 4 4\n0.1350 4 4\n0.1360 4 4\n0.1370 4 4\n0.1380 4 4\n0.1390 4 4\n0.1400 4 4\n0.1410 4 4\n0.1420 4 4\n0.1430 4 4\n0.1440 4 4\n0.1450 4 4\n0.1460 4 4\n0.1470 4 4\n0.1480 4 4\n0.1490 4 4\n0.1500 4 4\n0.1510 4 4\n0.1520 4 4\n0.1530 4 4\n0.1540 4 4\n0.1550 4 4\n0.1560 4 4\n0.1570 4 4\n0.1580 4 4\n0.1590 4 4\n0.1600 4 4\n0.1610 4 4\n0.1620 4 4\n0.1630 4 4\n0.1640 4 4\n0.1650 4 4\n0.1660 4 4\n0.1670 4 4\n0.1680 4 4\n0.1690 4 4\n0.1700 4 4\n0.1710 4 4\n0.1720 4 4\n0.1730 4 4\n0.1740 4 4\n0.1750 4 4\n0.1760 4 4\n0.1770 4 4\n0.1780 4 4\n0.1790 4 4\n0.1800 4 4\n0.1810 4 4\n0.1820 4 4\n0.1830 4 4\n0.1840 4 4\n0.1850 4 4\n0.1860 4 4\n0.1870 4 4\n0.1880 4 4\n0.1890 4 4\n0.1900 4 4\n0.1910 4 4\n0.1920 4 4\n0.1930 4 4\n0.1940 4 4\n0.1950 4 4\n0.1960 4 4\n0.1970 4 4\n0.1980 4 4\n0.1990 4 4\n</code></pre>\n" }, { "answer_id": 260571, "score": 2, "text": "<p>I am letting $A = H_n - \\lfloor H_n \\rfloor$ and computing it in a dumb way, and then computing $1/nA$ when $A \\lt 1/n $. For $n=83$ I get $1/nA \\approx 5.825$, which is the largest such value I find so far, and which means 83 is one of three values of $n$ which are allowed when $\\epsilon$ is 0.3. A smarter way to compute the values of $n$ for which $1/nA$ is worth computing is to use $\\log(n+1/2) + \\gamma$ for just those $n$ for which the expression is very near an integer, for $H_n$ is within $1/24n^2$ of this value. $\\log(1/nA )/\\log(n)$ then serves as an upper bound on $\\epsilon$.</p>\n\n<p>My calculations show $1 \\lt 1/nA \\lt 2$ much of the time, which means $\\epsilon \\lt 1/\\log n$ for admissible $n$. I expect the conjecture to hold and when proved to reveal some of the nature of Euler's $\\gamma$.</p>\n\n<p>Gerhard \"Still Computing Oh So Slowly\" Paseman, 2017.01.25.</p>\n" }, { "answer_id": 260669, "score": 4, "text": "<p><strong><em>This is a comment at the comments of Gerhard \"Still Computing Oh So Slowly\" Paseman's answer, giving just indexes n for more record-holders.</em></strong> </p>\n\n<p>Let $\\small h_n$ denote the <em>n</em>'th harmonic number, $\\small A_n=\\{h_n\\}$ its fractional part. The sequence of $A_n$ has a remarkable shape like a sawtooth-curve with increasing wavelength when <em>n</em> is increased and with sharp local minima - a shape which can be exploited when we seek for possible <em>n</em> to be included in <strong><em>M</em></strong>.<br>\nI use $\\small f(n) = w_n = \\{h_n\\} \\cdot n$ and $\\small f(n,\\varepsilon)=w_n \\cdot n^\\varepsilon $ with the OP's condition rewritten as $\\small f(n,\\varepsilon)&lt;1 $ as criterion for the inclusion of <em>n</em> into the set <strong><em>M</em></strong> . Of course the cardinality of <strong><em>M</em></strong> is limited by an upper bound $ \\small n \\le N$ with some <strong><em>N</em></strong> that can numerically be handled, so actually we should write explicitely $\\small |M(\\varepsilon,N)| $ instead of <strong><em>M</em></strong> only. I could manage to use $\\small N \\approx e^{2000}$ with the help of Pari/GP. </p>\n\n<p>The local minima of $\\small A_n$ occur near $\\small x_k=e^{k-\\gamma}, k \\in \\mathbb N^+$ and the <em>n</em> to be tested is one of the next integers enclosing the $\\small x_k$ - so this exact <em>n</em> must be empirically be determined. (Note that my <em>n</em> are the <em>n+1</em> used in the comments at the OP denoting the previous <em>high</em> value of <strong><em>A</em></strong> ) </p>\n\n<p>The harmonic numbers $\\small h_n$ can be computed in Pari/GP using <code>h(n) = psi(1+n) + Euler</code> ; however, this seems to be limited to something like $ \\small n \\lt e^{600}$ and so I had to introduce the Euler-McLaurin-formula for the larger <em>n</em> and implemented the switch from one method to the other at <code>n = 1e50</code> </p>\n\n<p><hr>\nThe following table shows the first <em>40</em> entries of my <em>2000</em>-row table containing $\\small n \\approx 10^{840}$ which was possible to compute using Pari/GP in just a couple of seconds. </p>\n\n<p>The basic option according to the focus in the OP's question is to test $\\small f(n,\\varepsilon) = w_n \\cdot n^\\varepsilon \\lt 1$. For an example with $\\small \\varepsilon=0.1$ see the 6'th column and the 7'th column allowing to sum to the cardinality of <strong><em>M</em></strong> where we find $\\small | M(0.1,2000) | = 7$ with $\\small n \\le N \\approx e^{2000}$ </p>\n\n<p>A second, much nicer, option is to define the function $\\small e(n) = -\\log_n(w_n) $ and test $\\small e(n) \\gt \\varepsilon$ whether to include this <em>n</em> or not. This is simply possible when looking into the <em>5</em>'th column and simply compare.<br>\nThis latter method allows to compute the cardinalities of $\\small M(\\varepsilon)$ for arbitrary $\\varepsilon$ really fast, for instance to create informative scatter- or lineplots, when first a list for $e(n)$ of length $\\small N$ is created and then the successful comparisions with the intended $\\varepsilon$ are summed to determine the cardinality. </p>\n\n<pre><code> n | h_n | A=frac(h_n) | w=A*n | e(n) | w*n^0.1|in M?\n -----------------+--------+-------------+----------+-----------+--------+---\n 1 1.00000 0.00000 0.00000 1.000000 0.00000 1\n 4 2.08333 0.0833333 0.333333 0.792481 0.382899 1\n 11 3.01988 0.0198773 0.218651 0.634006 0.277901 1\n 31 4.02725 0.0272452 0.844601 0.0491822 1.19066 .\n 83 5.00207 0.00206827 0.171667 0.398793 0.267051 1\n 227 6.00437 0.00436671 0.991243 0.00162136 1.70523 .\n 616 7.00127 0.00127410 0.784844 0.0377178 1.49191 .\n 1674 8.00049 0.000485572 0.812848 0.0279149 1.70759 .\n 4550 9.00021 0.000208063 0.946686 0.00650460 2.19790 .\n 12367 10.0000 0.0000430083 0.531883 0.0670005 1.36472 .\n 33617 11.0000 0.0000177086 0.595311 0.0497632 1.68811 .\n 91380 12.0000 0.00000305167 0.278861 0.111798 0.873923 1\n 248397 13.0000 0.00000122948 0.305399 0.0954806 1.05775 .\n 675214 14.0000 1.36205E-7 0.919678 0.00623806 3.52030 .\n 1835421 15.0000 3.78268E-7 0.694281 0.0252988 2.93703 .\n 4989191 16.0000 9.54538E-7 0.476237 0.0481002 2.22652 .\n 13562027 17.0000 1.48499E-8 0.201395 0.0975770 1.04059 .\n 36865412 18.0000 3.71993E-9 0.137137 0.114033 0.783098 1\n 100210581 19.0000 9.73330E-9 0.975380 0.00135314 6.15552 .\n 272400600 20.0000 1.61744E-9 0.440592 0.0421997 3.07297 .\n 740461601 21.0000 4.01333E-9 0.297172 0.0594162 2.29065 .\n 2012783315 22.0000 1.38447E-10 0.278664 0.0596444 2.37389 .\n 5471312310 23.0000 1.97920E-11 0.108288 0.0991384 1.01951 .\n 14872568831 24.0000 2.27220E-11 0.337935 0.0463183 3.51618 .\n 40427833596 25.0000 6.07937E-12 0.245776 0.0574601 2.82623 .\n 109894245429 26.0000 7.60776E-12 0.836049 0.00704359 10.6250 .\n 298723530401 27.0000 1.82203E-12 0.544283 0.0230213 7.64454 .\n 812014744422 28.0000 5.52830E-13 0.448906 0.0292071 6.96806 .\n 2207284924203 29.0000 1.00870E-13 0.222650 0.0528504 3.81951 .\n 6000022499693 30.0000 2.16954E-14 0.130173 0.0692963 2.46795 .\n 16309752131262 31.0000 3.65111E-14 0.595487 0.0170391 12.4772 .\n 44334502845080 32.0000 1.81005E-15 0.080247 0.0802804 1.85827 .\n 120513673457548 33.0000 4.59281E-15 0.553496 0.0182434 14.1651 .\n 327590128640500 34.0000 2.31992E-15 0.759983 0.00821173 21.4950 .\n 890482293866031 35.0000 2.71425E-17 0.024169 0.108145 0.755506 1\n 2420581837980561 36.0000 2.55560E-16 0.618603 0.0135588 21.3700 .\n 6579823624480555 37.0000 1.20561E-17 0.079326 0.0695767 3.02860 .\n 17885814992891026 38.0000 4.26642E-18 0.076308 0.0687541 3.21976 .\n 48618685882356024 39.0000 7.23781E-19 0.035189 0.0871101 1.64093 .\n132159290357566703 40.0000 2.02186E-18 0.267208 0.0334763 13.7708 .\n</code></pre>\n\n<p>The set $\\small M(0.1,e^{40})$ (represented by <em>40</em> rows) is here the set of all $n$ where $\\small f(n,0.1)=w_n \\cdot n^{0.1} \\lt 1$ . The number of such entries for $\\small n=132159290357566703 \\approx 10^{17}$ is here <em>7</em> . </p>\n\n<p>All numerical tests which I've done indicate that the cardinality of $\\small M(\\varepsilon \\gt 0,\\infty)$ is finite and roughly reciprocal to $\\small \\varepsilon$ and only if $\\varepsilon=0$ is surely infinite. </p>\n\n<p><hr></p>\n\n<h1>Pari/GP tools</h1>\n\n<p>This is the Pari/GP-program which I used. </p>\n\n<p>Using the functions</p>\n\n<pre><code> h(n) = Euler+if(n&lt;1e50, psi(1+n),log(n)+1/2/n-1/12/n^2+1/120/n^4-1/252/n^6+1/240/n^8)\n A(n) = if(n==1,return(0)); frac(h(n))\n w(n) = A(n)*n \n e(n) = if(n==1,return(1)); -log(w(n))/log(n)\n</code></pre>\n\n<p>The following needs only two steps to find the <em>n</em> with the next local minimum when called with index $\\small k \\in \\mathbb N$ : </p>\n\n<pre><code> {find_n(k)=local(n1,n2,a1,a2);\n n1=floor(exp(k-Euler)); n2=n1+1; a1 = A(n1);a2 = A(n2);\n if(a1&lt;a2,return(n1),return(n2)); \n } \n\n\n\n \\\\ create that list one time, then evaluate cardinality for various eps\n \\\\ using that same list\n {makeList(listlen=40)= local(list,n,w1,logn);\n list=matrix(listlen,4);\n list[1,]=[Euler,0,0,0];\n for(k=2,listlen,\n n=find_n(k); logn=log(n);\n list[k,]=[logn+Euler,n,w1=w(n),-log(w1)/logn];\n );\n return(list); }\n\n \\\\ compute cardinality with some eps\n cardM(eps,list) = sum(k=1,#list[,1],list[k,4]&gt;eps) \n\n \\\\ apply, note: for long lists we need high internal precision\n list = makeList(40) \\\\ the max N is here about e^40\n print(cardM(0.1,list) )\n</code></pre>\n\n<hr>\n\n<p><strong><em>[table 2]</em></strong>: This are sample data for the OP's plot of the cardinality of $\\small M(\\varepsilon)$ by the argument $\\varepsilon$ where $\\small N \\approx \\exp(250) \\approx 10^{108}$ : </p>\n\n<pre><code> eps | c=#M | r=c*eps | c= card(M) for n&lt;= N approx 10^108\n ------+------------------\n 0.01 104 1.040\n 0.02 55 1.100\n 0.03 36 1.080\n 0.04 29 1.160\n 0.05 21 1.050\n 0.06 16 0.960\n 0.07 12 0.840\n 0.08 12 0.960\n 0.09 10 0.900\n 0.10 7 0.700\n 0.11 6 0.660\n 0.12 4 0.480\n 0.13 4 0.520\n 0.14 4 0.560\n 0.15 4 0.600\n 0.16 4 0.640\n 0.17 4 0.680\n 0.18 4 0.720\n 0.19 4 0.760\n 0.20 4 0.800\n</code></pre>\n\n<p><hr></p>\n\n<h1>Pictures</h1>\n\n<p><strong><em>[Picture 1]</em></strong> : I've extended the search-space for $n$ to the range $\\small 1 \\ldots e^{1000} \\approx 10^{434} $ and to check sanity to the range $\\small 1 \\ldots e^{2000} \\approx 10^{868} $. For epsilons $\\varepsilon$ from $0.001$ in $200$ steps up to $0.2$ I made the following graph:<br>\n<a href=\"https://i.stack.imgur.com/DAz5g.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/DAz5g.png\" alt=\"picture\"></a> </p>\n\n<p>Remarks: for the very small epsilon the increase of the search-space gives slightly higher results, which also illustrates, that for \"larger\" epsilon the cardinality of $\\small M(\\varepsilon)$ is finite</p>\n\n<p><hr>\n<strong><em>[Picture 2]</em></strong>: Indicates uniformity of the $\\small f(n) =w_n = frac(h_n) \\cdot n^1$ at the <em>n</em>, where $\\small w_n $ has a local minimum (on request of @GerhardPaseman):<br>\n<a href=\"https://i.stack.imgur.com/18Az3.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/18Az3.png\" alt=\"picture 2\"></a> </p>\n\n<p><strong><em>[Picture 3]</em></strong>:It is also convenient to show a rescaling of the $\\small f(n)$ so that we can immediately determine the cardinalities $\\small |M(\\varepsilon)|$ just by counting the number of dots $\\small e(n)$ above $\\small \\varepsilon$. The derivation of the formula is </p>\n\n<p>$$ \\small{ \\begin{array}{lll}\n A(n) &amp;\\lt &amp; {1\\over n^1\\cdot n^\\varepsilon} &amp; \\text{ from OP}\\\\\n f(n) \\cdot n^\\varepsilon &amp;\\lt &amp; 1\\\\\n \\ln(f(n)) + \\varepsilon \\cdot \\ln(n) &amp;\\lt &amp; 0 \\\\\n {\\ln(f(n)) \\over \\ln(n)} + \\varepsilon &amp;\\lt &amp; 0 \\\\\n \\varepsilon &amp; \\lt&amp; -\\ln_n(f(n)) \n\\end{array} }$$</p>\n\n<p>and we simply count, how many dots in the picture show $\\small e(n) = - \\ln_n(f(n)) \\gt \\varepsilon$ </p>\n\n<p><a href=\"https://i.stack.imgur.com/QFKcS.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/QFKcS.png\" alt=\"picture\"></a> </p>\n\n<p><hr></p>\n\n<h1>Data for cardinalities-plot</h1>\n\n<pre><code>epsilon |M(eps,N)||M(eps,N)| \n N~e^1000 N~e^2000\n-------------------------------\n0.0000 1000 2000\n0.0010 633 862\n0.0020 430 482\n0.0030 315 330\n0.0040 258 264\n0.0050 204 205\n0.0060 171 171\n0.0070 151 151\n0.0080 135 135\n0.0090 120 120\n0.0100 109 109\n0.0110 100 100\n0.0120 98 98\n0.0130 89 89\n0.0140 84 84\n0.0150 79 79\n0.0160 73 73\n0.0170 69 69\n0.0180 62 62\n0.0190 57 57\n0.0200 55 55\n0.0210 50 50\n0.0220 47 47\n0.0230 44 44\n0.0240 42 42\n0.0250 41 41\n0.0260 39 39\n0.0270 39 39\n0.0280 37 37\n0.0290 37 37\n0.0300 36 36\n0.0310 36 36\n0.0320 36 36\n0.0330 36 36\n0.0340 34 34\n0.0350 34 34\n0.0360 34 34\n0.0370 33 33\n0.0380 32 32\n0.0390 31 31\n0.0400 29 29\n0.0410 29 29\n0.0420 29 29\n0.0430 26 26\n0.0440 26 26\n0.0450 25 25\n0.0460 25 25\n0.0470 24 24\n0.0480 24 24\n0.0490 23 23\n0.0500 21 21\n0.0510 21 21\n0.0520 21 21\n0.0530 20 20\n0.0540 20 20\n0.0550 20 20\n0.0560 20 20\n0.0570 19 19\n0.0580 18 18\n0.0590 18 18\n0.0600 16 16\n0.0610 16 16\n0.0620 16 16\n0.0630 16 16\n0.0640 16 16\n0.0650 16 16\n0.0660 16 16\n0.0670 16 16\n0.0680 15 15\n0.0690 14 14\n0.0700 12 12\n0.0710 12 12\n0.0720 12 12\n0.0730 12 12\n0.0740 12 12\n0.0750 12 12\n0.0760 12 12\n0.0770 12 12\n0.0780 12 12\n0.0790 12 12\n0.0800 12 12\n0.0810 11 11\n0.0820 11 11\n0.0830 11 11\n0.0840 11 11\n0.0850 11 11\n0.0860 11 11\n0.0870 11 11\n0.0880 10 10\n0.0890 10 10\n0.0900 10 10\n0.0910 10 10\n0.0920 10 10\n0.0930 10 10\n0.0940 10 10\n0.0950 10 10\n0.0960 9 9\n0.0970 9 9\n0.0980 8 8\n0.0990 8 8\n0.1000 7 7\n0.1010 7 7\n0.1020 7 7\n0.1030 7 7\n0.1040 7 7\n0.1050 7 7\n0.1060 7 7\n0.1070 7 7\n0.1080 7 7\n0.1090 6 6\n0.1100 6 6\n0.1110 6 6\n0.1120 5 5\n0.1130 5 5\n0.1140 5 5\n0.1150 4 4\n0.1160 4 4\n0.1170 4 4\n0.1180 4 4\n0.1190 4 4\n0.1200 4 4\n0.1210 4 4\n0.1220 4 4\n0.1230 4 4\n0.1240 4 4\n0.1250 4 4\n0.1260 4 4\n0.1270 4 4\n0.1280 4 4\n0.1290 4 4\n0.1300 4 4\n0.1310 4 4\n0.1320 4 4\n0.1330 4 4\n0.1340 4 4\n0.1350 4 4\n0.1360 4 4\n0.1370 4 4\n0.1380 4 4\n0.1390 4 4\n0.1400 4 4\n0.1410 4 4\n0.1420 4 4\n0.1430 4 4\n0.1440 4 4\n0.1450 4 4\n0.1460 4 4\n0.1470 4 4\n0.1480 4 4\n0.1490 4 4\n0.1500 4 4\n0.1510 4 4\n0.1520 4 4\n0.1530 4 4\n0.1540 4 4\n0.1550 4 4\n0.1560 4 4\n0.1570 4 4\n0.1580 4 4\n0.1590 4 4\n0.1600 4 4\n0.1610 4 4\n0.1620 4 4\n0.1630 4 4\n0.1640 4 4\n0.1650 4 4\n0.1660 4 4\n0.1670 4 4\n0.1680 4 4\n0.1690 4 4\n0.1700 4 4\n0.1710 4 4\n0.1720 4 4\n0.1730 4 4\n0.1740 4 4\n0.1750 4 4\n0.1760 4 4\n0.1770 4 4\n0.1780 4 4\n0.1790 4 4\n0.1800 4 4\n0.1810 4 4\n0.1820 4 4\n0.1830 4 4\n0.1840 4 4\n0.1850 4 4\n0.1860 4 4\n0.1870 4 4\n0.1880 4 4\n0.1890 4 4\n0.1900 4 4\n0.1910 4 4\n0.1920 4 4\n0.1930 4 4\n0.1940 4 4\n0.1950 4 4\n0.1960 4 4\n0.1970 4 4\n0.1980 4 4\n0.1990 4 4\n</code></pre>\n" }, { "answer_id": 260571, "score": 2, "text": "<p>I am letting $A = H_n - \\lfloor H_n \\rfloor$ and computing it in a dumb way, and then computing $1/nA$ when $A \\lt 1/n $. For $n=83$ I get $1/nA \\approx 5.825$, which is the largest such value I find so far, and which means 83 is one of three values of $n$ which are allowed when $\\epsilon$ is 0.3. A smarter way to compute the values of $n$ for which $1/nA$ is worth computing is to use $\\log(n+1/2) + \\gamma$ for just those $n$ for which the expression is very near an integer, for $H_n$ is within $1/24n^2$ of this value. $\\log(1/nA )/\\log(n)$ then serves as an upper bound on $\\epsilon$.</p>\n\n<p>My calculations show $1 \\lt 1/nA \\lt 2$ much of the time, which means $\\epsilon \\lt 1/\\log n$ for admissible $n$. I expect the conjecture to hold and when proved to reveal some of the nature of Euler's $\\gamma$.</p>\n\n<p>Gerhard \"Still Computing Oh So Slowly\" Paseman, 2017.01.25.</p>\n" } ]
Question: Conjecture about harmonic numbers Answer 1: <p><strong><em>This is a comment at the comments of Gerhard "Still Computing Oh So Slowly" Paseman's answer, giving just indexes n for more record-holders.</em></strong> </p> <p>Let $\small h_n$ denote the <em>n</em>'th harmonic number, $\small A_n=\{h_n\}$ its fractional part. The sequence of $A_n$ has a remarkable shape like a sawtooth-curve with increasing wavelength when <em>n</em> is increased and with sharp local minima - a shape which can be exploited when we seek for possible <em>n</em> to be included in <strong><em>M</em></strong>.<br> I use $\small f(n) = w_n = \{h_n\} \cdot n$ and $\small f(n,\varepsilon)=w_n \cdot n^\varepsilon $ with the OP's condition rewritten as $\small f(n,\varepsilon)&lt;1 $ as criterion for the inclusion of <em>n</em> into the set <strong><em>M</em></strong> . Of course the cardinality of <strong><em>M</em></strong> is limited by an upper bound $ \small n \le N$ with some <strong><em>N</em></strong> that can numerically be handled, so actually we should write explicitely $\small |M(\varepsilon,N)| $ instead of <strong><em>M</em></strong> only. I could manage to use $\small N \approx e^{2000}$ with the help of Pari/GP. </p> <p>The local minima of $\small A_n$ occur near $\small x_k=e^{k-\gamma}, k \in \mathbb N^+$ and the <em>n</em> to be tested is one of the next integers enclosing the $\small x_k$ - so this exact <em>n</em> must be empirically be determined. (Note that my <em>n</em> are the <em>n+1</em> used in the comments at the OP denoting the previous <em>high</em> value of <strong><em>A</em></strong> ) </p> <p>The harmonic numbers $\small h_n$ can be computed in Pari/GP using <code>h(n) = psi(1+n) + Euler</code> ; however, this seems to be limited to something like $ \small n \lt e^{600}$ and so I had to introduce the Euler-McLaurin-formula for the larger <em>n</em> and implemented the switch from one method to the other at <code>n = 1e50</code> </p> <p><hr> The following table shows the first <em>40</em> entries of my <em>2000</em>-row table containing $\small n \approx 10^{840}$ which was possible to compute using Pari/GP in just a couple of seconds. </p> <p>The basic option according to the focus in the OP's question is to test $\small f(n,\varepsilon) = w_n \cdot n^\varepsilon \lt 1$. For an example with $\small \varepsilon=0.1$ see the 6'th column and the 7'th column allowing to sum to the cardinality of <strong><em>M</em></strong> where we find $\small | M(0.1,2000) | = 7$ with $\small n \le N \approx e^{2000}$ </p> <p>A second, much nicer, option is to define the function $\small e(n) = -\log_n(w_n) $ and test $\small e(n) \gt \varepsilon$ whether to include this <em>n</em> or not. This is simply possible when looking into the <em>5</em>'th column and simply compare.<br> This latter method allows to compute the cardinalities of $\small M(\varepsilon)$ for arbitrary $\varepsilon$ really fast, for instance to create informative scatter- or lineplots, when first a list for $e(n)$ of length $\small N$ is created and then the successful comparisions with the intended $\varepsilon$ are summed to determine the cardinality. </p> <pre><code> n | h_n | A=frac(h_n) | w=A*n | e(n) | w*n^0.1|in M? -----------------+--------+-------------+----------+-----------+--------+--- 1 1.00000 0.00000 0.00000 1.000000 0.00000 1 4 2.08333 0.0833333 0.333333 0.792481 0.382899 1 11 3.01988 0.0198773 0.218651 0.634006 0.277901 1 31 4.02725 0.0272452 0.844601 0.0491822 1.19066 . 83 5.00207 0.00206827 0.171667 0.398793 0.267051 1 227 6.00437 0.00436671 0.991243 0.00162136 1.70523 . 616 7.00127 0.00127410 0.784844 0.0377178 1.49191 . 1674 8.00049 0.000485572 0.812848 0.0279149 1.70759 . 4550 9.00021 0.000208063 0.946686 0.00650460 2.19790 . 12367 10.0000 0.0000430083 0.531883 0.0670005 1.36472 . 33617 11.0000 0.0000177086 0.595311 0.0497632 1.68811 . 91380 12.0000 0.00000305167 0.278861 0.111798 0.873923 1 248397 13.0000 0.00000122948 0.305399 0.0954806 1.05775 . 675214 14.0000 1.36205E-7 0.919678 0.00623806 3.52030 . 1835421 15.0000 3.78268E-7 0.694281 0.0252988 2.93703 . 4989191 16.0000 9.54538E-7 0.476237 0.0481002 2.22652 . 13562027 17.0000 1.48499E-8 0.201395 0.0975770 1.04059 . 36865412 18.0000 3.71993E-9 0.137137 0.114033 0.783098 1 100210581 19.0000 9.73330E-9 0.975380 0.00135314 6.15552 . 272400600 20.0000 1.61744E-9 0.440592 0.0421997 3.07297 . 740461601 21.0000 4.01333E-9 0.297172 0.0594162 2.29065 . 2012783315 22.0000 1.38447E-10 0.278664 0.0596444 2.37389 . 5471312310 23.0000 1.97920E-11 0.108288 0.0991384 1.01951 . 14872568831 24.0000 2.27220E-11 0.337935 0.0463183 3.51618 . 40427833596 25.0000 6.07937E-12 0.245776 0.0574601 2.82623 . 109894245429 26.0000 7.60776E-12 0.836049 0.00704359 10.6250 . 298723530401 27.0000 1.82203E-12 0.544283 0.0230213 7.64454 . 812014744422 28.0000 5.52830E-13 0.448906 0.0292071 6.96806 . 2207284924203 29.0000 1.00870E-13 0.222650 0.0528504 3.81951 . 6000022499693 30.0000 2.16954E-14 0.130173 0.0692963 2.46795 . 16309752131262 31.0000 3.65111E-14 0.595487 0.0170391 12.4772 . 44334502845080 32.0000 1.81005E-15 0.080247 0.0802804 1.85827 . 120513673457548 33.0000 4.59281E-15 0.553496 0.0182434 14.1651 . 327590128640500 34.0000 2.31992E-15 0.759983 0.00821173 21.4950 . 890482293866031 35.0000 2.71425E-17 0.024169 0.108145 0.755506 1 2420581837980561 36.0000 2.55560E-16 0.618603 0.0135588 21.3700 . 6579823624480555 37.0000 1.20561E-17 0.079326 0.0695767 3.02860 . 17885814992891026 38.0000 4.26642E-18 0.076308 0.0687541 3.21976 . 48618685882356024 39.0000 7.23781E-19 0.035189 0.0871101 1.64093 . 132159290357566703 40.0000 2.02186E-18 0.267208 0.0334763 13.7708 . </code></pre> <p>The set $\small M(0.1,e^{40})$ (represented by <em>40</em> rows) is here the set of all $n$ where $\small f(n,0.1)=w_n \cdot n^{0.1} \lt 1$ . The number of such entries for $\small n=132159290357566703 \approx 10^{17}$ is here <em>7</em> . </p> <p>All numerical tests which I've done indicate that the cardinality of $\small M(\varepsilon \gt 0,\infty)$ is finite and roughly reciprocal to $\small \varepsilon$ and only if $\varepsilon=0$ is surely infinite. </p> <p><hr></p> <h1>Pari/GP tools</h1> <p>This is the Pari/GP-program which I used. </p> <p>Using the functions</p> <pre><code> h(n) = Euler+if(n&lt;1e50, psi(1+n),log(n)+1/2/n-1/12/n^2+1/120/n^4-1/252/n^6+1/240/n^8) A(n) = if(n==1,return(0)); frac(h(n)) w(n) = A(n)*n e(n) = if(n==1,return(1)); -log(w(n))/log(n) </code></pre> <p>The following needs only two steps to find the <em>n</em> with the next local minimum when called with index $\small k \in \mathbb N$ : </p> <pre><code> {find_n(k)=local(n1,n2,a1,a2); n1=floor(exp(k-Euler)); n2=n1+1; a1 = A(n1);a2 = A(n2); if(a1&lt;a2,return(n1),return(n2)); } \\ create that list one time, then evaluate cardinality for various eps \\ using that same list {makeList(listlen=40)= local(list,n,w1,logn); list=matrix(listlen,4); list[1,]=[Euler,0,0,0]; for(k=2,listlen, n=find_n(k); logn=log(n); list[k,]=[logn+Euler,n,w1=w(n),-log(w1)/logn]; ); return(list); } \\ compute cardinality with some eps cardM(eps,list) = sum(k=1,#list[,1],list[k,4]&gt;eps) \\ apply, note: for long lists we need high internal precision list = makeList(40) \\ the max N is here about e^40 print(cardM(0.1,list) ) </code></pre> <hr> <p><strong><em>[table 2]</em></strong>: This are sample data for the OP's plot of the cardinality of $\small M(\varepsilon)$ by the argument $\varepsilon$ where $\small N \approx \exp(250) \approx 10^{108}$ : </p> <pre><code> eps | c=#M | r=c*eps | c= card(M) for n&lt;= N approx 10^108 ------+------------------ 0.01 104 1.040 0.02 55 1.100 0.03 36 1.080 0.04 29 1.160 0.05 21 1.050 0.06 16 0.960 0.07 12 0.840 0.08 12 0.960 0.09 10 0.900 0.10 7 0.700 0.11 6 0.660 0.12 4 0.480 0.13 4 0.520 0.14 4 0.560 0.15 4 0.600 0.16 4 0.640 0.17 4 0.680 0.18 4 0.720 0.19 4 0.760 0.20 4 0.800 </code></pre> <p><hr></p> <h1>Pictures</h1> <p><strong><em>[Picture 1]</em></strong> : I've extended the search-space for $n$ to the range $\small 1 \ldots e^{1000} \approx 10^{434} $ and to check sanity to the range $\small 1 \ldots e^{2000} \approx 10^{868} $. For epsilons $\varepsilon$ from $0.001$ in $200$ steps up to $0.2$ I made the following graph:<br> <a href="https://i.stack.imgur.com/DAz5g.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/DAz5g.png" alt="picture"></a> </p> <p>Remarks: for the very small epsilon the increase of the search-space gives slightly higher results, which also illustrates, that for "larger" epsilon the cardinality of $\small M(\varepsilon)$ is finite</p> <p><hr> <strong><em>[Picture 2]</em></strong>: Indicates uniformity of the $\small f(n) =w_n = frac(h_n) \cdot n^1$ at the <em>n</em>, where $\small w_n $ has a local minimum (on request of @GerhardPaseman):<br> <a href="https://i.stack.imgur.com/18Az3.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/18Az3.png" alt="picture 2"></a> </p> <p><strong><em>[Picture 3]</em></strong>:It is also convenient to show a rescaling of the $\small f(n)$ so that we can immediately determine the cardinalities $\small |M(\varepsilon)|$ just by counting the number of dots $\small e(n)$ above $\small \varepsilon$. The derivation of the formula is </p> <p>$$ \small{ \begin{array}{lll} A(n) &amp;\lt &amp; {1\over n^1\cdot n^\varepsilon} &amp; \text{ from OP}\\ f(n) \cdot n^\varepsilon &amp;\lt &amp; 1\\ \ln(f(n)) + \varepsilon \cdot \ln(n) &amp;\lt &amp; 0 \\ {\ln(f(n)) \over \ln(n)} + \varepsilon &amp;\lt &amp; 0 \\ \varepsilon &amp; \lt&amp; -\ln_n(f(n)) \end{array} }$$</p> <p>and we simply count, how many dots in the picture show $\small e(n) = - \ln_n(f(n)) \gt \varepsilon$ </p> <p><a href="https://i.stack.imgur.com/QFKcS.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/QFKcS.png" alt="picture"></a> </p> <p><hr></p> <h1>Data for cardinalities-plot</h1> <pre><code>epsilon |M(eps,N)||M(eps,N)| N~e^1000 N~e^2000 ------------------------------- 0.0000 1000 2000 0.0010 633 862 0.0020 430 482 0.0030 315 330 0.0040 258 264 0.0050 204 205 0.0060 171 171 0.0070 151 151 0.0080 135 135 0.0090 120 120 0.0100 109 109 0.0110 100 100 0.0120 98 98 0.0130 89 89 0.0140 84 84 0.0150 79 79 0.0160 73 73 0.0170 69 69 0.0180 62 62 0.0190 57 57 0.0200 55 55 0.0210 50 50 0.0220 47 47 0.0230 44 44 0.0240 42 42 0.0250 41 41 0.0260 39 39 0.0270 39 39 0.0280 37 37 0.0290 37 37 0.0300 36 36 0.0310 36 36 0.0320 36 36 0.0330 36 36 0.0340 34 34 0.0350 34 34 0.0360 34 34 0.0370 33 33 0.0380 32 32 0.0390 31 31 0.0400 29 29 0.0410 29 29 0.0420 29 29 0.0430 26 26 0.0440 26 26 0.0450 25 25 0.0460 25 25 0.0470 24 24 0.0480 24 24 0.0490 23 23 0.0500 21 21 0.0510 21 21 0.0520 21 21 0.0530 20 20 0.0540 20 20 0.0550 20 20 0.0560 20 20 0.0570 19 19 0.0580 18 18 0.0590 18 18 0.0600 16 16 0.0610 16 16 0.0620 16 16 0.0630 16 16 0.0640 16 16 0.0650 16 16 0.0660 16 16 0.0670 16 16 0.0680 15 15 0.0690 14 14 0.0700 12 12 0.0710 12 12 0.0720 12 12 0.0730 12 12 0.0740 12 12 0.0750 12 12 0.0760 12 12 0.0770 12 12 0.0780 12 12 0.0790 12 12 0.0800 12 12 0.0810 11 11 0.0820 11 11 0.0830 11 11 0.0840 11 11 0.0850 11 11 0.0860 11 11 0.0870 11 11 0.0880 10 10 0.0890 10 10 0.0900 10 10 0.0910 10 10 0.0920 10 10 0.0930 10 10 0.0940 10 10 0.0950 10 10 0.0960 9 9 0.0970 9 9 0.0980 8 8 0.0990 8 8 0.1000 7 7 0.1010 7 7 0.1020 7 7 0.1030 7 7 0.1040 7 7 0.1050 7 7 0.1060 7 7 0.1070 7 7 0.1080 7 7 0.1090 6 6 0.1100 6 6 0.1110 6 6 0.1120 5 5 0.1130 5 5 0.1140 5 5 0.1150 4 4 0.1160 4 4 0.1170 4 4 0.1180 4 4 0.1190 4 4 0.1200 4 4 0.1210 4 4 0.1220 4 4 0.1230 4 4 0.1240 4 4 0.1250 4 4 0.1260 4 4 0.1270 4 4 0.1280 4 4 0.1290 4 4 0.1300 4 4 0.1310 4 4 0.1320 4 4 0.1330 4 4 0.1340 4 4 0.1350 4 4 0.1360 4 4 0.1370 4 4 0.1380 4 4 0.1390 4 4 0.1400 4 4 0.1410 4 4 0.1420 4 4 0.1430 4 4 0.1440 4 4 0.1450 4 4 0.1460 4 4 0.1470 4 4 0.1480 4 4 0.1490 4 4 0.1500 4 4 0.1510 4 4 0.1520 4 4 0.1530 4 4 0.1540 4 4 0.1550 4 4 0.1560 4 4 0.1570 4 4 0.1580 4 4 0.1590 4 4 0.1600 4 4 0.1610 4 4 0.1620 4 4 0.1630 4 4 0.1640 4 4 0.1650 4 4 0.1660 4 4 0.1670 4 4 0.1680 4 4 0.1690 4 4 0.1700 4 4 0.1710 4 4 0.1720 4 4 0.1730 4 4 0.1740 4 4 0.1750 4 4 0.1760 4 4 0.1770 4 4 0.1780 4 4 0.1790 4 4 0.1800 4 4 0.1810 4 4 0.1820 4 4 0.1830 4 4 0.1840 4 4 0.1850 4 4 0.1860 4 4 0.1870 4 4 0.1880 4 4 0.1890 4 4 0.1900 4 4 0.1910 4 4 0.1920 4 4 0.1930 4 4 0.1940 4 4 0.1950 4 4 0.1960 4 4 0.1970 4 4 0.1980 4 4 0.1990 4 4 </code></pre> Answer 2: <p>I am letting $A = H_n - \lfloor H_n \rfloor$ and computing it in a dumb way, and then computing $1/nA$ when $A \lt 1/n $. For $n=83$ I get $1/nA \approx 5.825$, which is the largest such value I find so far, and which means 83 is one of three values of $n$ which are allowed when $\epsilon$ is 0.3. A smarter way to compute the values of $n$ for which $1/nA$ is worth computing is to use $\log(n+1/2) + \gamma$ for just those $n$ for which the expression is very near an integer, for $H_n$ is within $1/24n^2$ of this value. $\log(1/nA )/\log(n)$ then serves as an upper bound on $\epsilon$.</p> <p>My calculations show $1 \lt 1/nA \lt 2$ much of the time, which means $\epsilon \lt 1/\log n$ for admissible $n$. I expect the conjecture to hold and when proved to reveal some of the nature of Euler's $\gamma$.</p> <p>Gerhard "Still Computing Oh So Slowly" Paseman, 2017.01.25.</p> Answer 3: <p><strong><em>This is a comment at the comments of Gerhard "Still Computing Oh So Slowly" Paseman's answer, giving just indexes n for more record-holders.</em></strong> </p> <p>Let $\small h_n$ denote the <em>n</em>'th harmonic number, $\small A_n=\{h_n\}$ its fractional part. The sequence of $A_n$ has a remarkable shape like a sawtooth-curve with increasing wavelength when <em>n</em> is increased and with sharp local minima - a shape which can be exploited when we seek for possible <em>n</em> to be included in <strong><em>M</em></strong>.<br> I use $\small f(n) = w_n = \{h_n\} \cdot n$ and $\small f(n,\varepsilon)=w_n \cdot n^\varepsilon $ with the OP's condition rewritten as $\small f(n,\varepsilon)&lt;1 $ as criterion for the inclusion of <em>n</em> into the set <strong><em>M</em></strong> . Of course the cardinality of <strong><em>M</em></strong> is limited by an upper bound $ \small n \le N$ with some <strong><em>N</em></strong> that can numerically be handled, so actually we should write explicitely $\small |M(\varepsilon,N)| $ instead of <strong><em>M</em></strong> only. I could manage to use $\small N \approx e^{2000}$ with the help of Pari/GP. </p> <p>The local minima of $\small A_n$ occur near $\small x_k=e^{k-\gamma}, k \in \mathbb N^+$ and the <em>n</em> to be tested is one of the next integers enclosing the $\small x_k$ - so this exact <em>n</em> must be empirically be determined. (Note that my <em>n</em> are the <em>n+1</em> used in the comments at the OP denoting the previous <em>high</em> value of <strong><em>A</em></strong> ) </p> <p>The harmonic numbers $\small h_n$ can be computed in Pari/GP using <code>h(n) = psi(1+n) + Euler</code> ; however, this seems to be limited to something like $ \small n \lt e^{600}$ and so I had to introduce the Euler-McLaurin-formula for the larger <em>n</em> and implemented the switch from one method to the other at <code>n = 1e50</code> </p> <p><hr> The following table shows the first <em>40</em> entries of my <em>2000</em>-row table containing $\small n \approx 10^{840}$ which was possible to compute using Pari/GP in just a couple of seconds. </p> <p>The basic option according to the focus in the OP's question is to test $\small f(n,\varepsilon) = w_n \cdot n^\varepsilon \lt 1$. For an example with $\small \varepsilon=0.1$ see the 6'th column and the 7'th column allowing to sum to the cardinality of <strong><em>M</em></strong> where we find $\small | M(0.1,2000) | = 7$ with $\small n \le N \approx e^{2000}$ </p> <p>A second, much nicer, option is to define the function $\small e(n) = -\log_n(w_n) $ and test $\small e(n) \gt \varepsilon$ whether to include this <em>n</em> or not. This is simply possible when looking into the <em>5</em>'th column and simply compare.<br> This latter method allows to compute the cardinalities of $\small M(\varepsilon)$ for arbitrary $\varepsilon$ really fast, for instance to create informative scatter- or lineplots, when first a list for $e(n)$ of length $\small N$ is created and then the successful comparisions with the intended $\varepsilon$ are summed to determine the cardinality. </p> <pre><code> n | h_n | A=frac(h_n) | w=A*n | e(n) | w*n^0.1|in M? -----------------+--------+-------------+----------+-----------+--------+--- 1 1.00000 0.00000 0.00000 1.000000 0.00000 1 4 2.08333 0.0833333 0.333333 0.792481 0.382899 1 11 3.01988 0.0198773 0.218651 0.634006 0.277901 1 31 4.02725 0.0272452 0.844601 0.0491822 1.19066 . 83 5.00207 0.00206827 0.171667 0.398793 0.267051 1 227 6.00437 0.00436671 0.991243 0.00162136 1.70523 . 616 7.00127 0.00127410 0.784844 0.0377178 1.49191 . 1674 8.00049 0.000485572 0.812848 0.0279149 1.70759 . 4550 9.00021 0.000208063 0.946686 0.00650460 2.19790 . 12367 10.0000 0.0000430083 0.531883 0.0670005 1.36472 . 33617 11.0000 0.0000177086 0.595311 0.0497632 1.68811 . 91380 12.0000 0.00000305167 0.278861 0.111798 0.873923 1 248397 13.0000 0.00000122948 0.305399 0.0954806 1.05775 . 675214 14.0000 1.36205E-7 0.919678 0.00623806 3.52030 . 1835421 15.0000 3.78268E-7 0.694281 0.0252988 2.93703 . 4989191 16.0000 9.54538E-7 0.476237 0.0481002 2.22652 . 13562027 17.0000 1.48499E-8 0.201395 0.0975770 1.04059 . 36865412 18.0000 3.71993E-9 0.137137 0.114033 0.783098 1 100210581 19.0000 9.73330E-9 0.975380 0.00135314 6.15552 . 272400600 20.0000 1.61744E-9 0.440592 0.0421997 3.07297 . 740461601 21.0000 4.01333E-9 0.297172 0.0594162 2.29065 . 2012783315 22.0000 1.38447E-10 0.278664 0.0596444 2.37389 . 5471312310 23.0000 1.97920E-11 0.108288 0.0991384 1.01951 . 14872568831 24.0000 2.27220E-11 0.337935 0.0463183 3.51618 . 40427833596 25.0000 6.07937E-12 0.245776 0.0574601 2.82623 . 109894245429 26.0000 7.60776E-12 0.836049 0.00704359 10.6250 . 298723530401 27.0000 1.82203E-12 0.544283 0.0230213 7.64454 . 812014744422 28.0000 5.52830E-13 0.448906 0.0292071 6.96806 . 2207284924203 29.0000 1.00870E-13 0.222650 0.0528504 3.81951 . 6000022499693 30.0000 2.16954E-14 0.130173 0.0692963 2.46795 . 16309752131262 31.0000 3.65111E-14 0.595487 0.0170391 12.4772 . 44334502845080 32.0000 1.81005E-15 0.080247 0.0802804 1.85827 . 120513673457548 33.0000 4.59281E-15 0.553496 0.0182434 14.1651 . 327590128640500 34.0000 2.31992E-15 0.759983 0.00821173 21.4950 . 890482293866031 35.0000 2.71425E-17 0.024169 0.108145 0.755506 1 2420581837980561 36.0000 2.55560E-16 0.618603 0.0135588 21.3700 . 6579823624480555 37.0000 1.20561E-17 0.079326 0.0695767 3.02860 . 17885814992891026 38.0000 4.26642E-18 0.076308 0.0687541 3.21976 . 48618685882356024 39.0000 7.23781E-19 0.035189 0.0871101 1.64093 . 132159290357566703 40.0000 2.02186E-18 0.267208 0.0334763 13.7708 . </code></pre> <p>The set $\small M(0.1,e^{40})$ (represented by <em>40</em> rows) is here the set of all $n$ where $\small f(n,0.1)=w_n \cdot n^{0.1} \lt 1$ . The number of such entries for $\small n=132159290357566703 \approx 10^{17}$ is here <em>7</em> . </p> <p>All numerical tests which I've done indicate that the cardinality of $\small M(\varepsilon \gt 0,\infty)$ is finite and roughly reciprocal to $\small \varepsilon$ and only if $\varepsilon=0$ is surely infinite. </p> <p><hr></p> <h1>Pari/GP tools</h1> <p>This is the Pari/GP-program which I used. </p> <p>Using the functions</p> <pre><code> h(n) = Euler+if(n&lt;1e50, psi(1+n),log(n)+1/2/n-1/12/n^2+1/120/n^4-1/252/n^6+1/240/n^8) A(n) = if(n==1,return(0)); frac(h(n)) w(n) = A(n)*n e(n) = if(n==1,return(1)); -log(w(n))/log(n) </code></pre> <p>The following needs only two steps to find the <em>n</em> with the next local minimum when called with index $\small k \in \mathbb N$ : </p> <pre><code> {find_n(k)=local(n1,n2,a1,a2); n1=floor(exp(k-Euler)); n2=n1+1; a1 = A(n1);a2 = A(n2); if(a1&lt;a2,return(n1),return(n2)); } \\ create that list one time, then evaluate cardinality for various eps \\ using that same list {makeList(listlen=40)= local(list,n,w1,logn); list=matrix(listlen,4); list[1,]=[Euler,0,0,0]; for(k=2,listlen, n=find_n(k); logn=log(n); list[k,]=[logn+Euler,n,w1=w(n),-log(w1)/logn]; ); return(list); } \\ compute cardinality with some eps cardM(eps,list) = sum(k=1,#list[,1],list[k,4]&gt;eps) \\ apply, note: for long lists we need high internal precision list = makeList(40) \\ the max N is here about e^40 print(cardM(0.1,list) ) </code></pre> <hr> <p><strong><em>[table 2]</em></strong>: This are sample data for the OP's plot of the cardinality of $\small M(\varepsilon)$ by the argument $\varepsilon$ where $\small N \approx \exp(250) \approx 10^{108}$ : </p> <pre><code> eps | c=#M | r=c*eps | c= card(M) for n&lt;= N approx 10^108 ------+------------------ 0.01 104 1.040 0.02 55 1.100 0.03 36 1.080 0.04 29 1.160 0.05 21 1.050 0.06 16 0.960 0.07 12 0.840 0.08 12 0.960 0.09 10 0.900 0.10 7 0.700 0.11 6 0.660 0.12 4 0.480 0.13 4 0.520 0.14 4 0.560 0.15 4 0.600 0.16 4 0.640 0.17 4 0.680 0.18 4 0.720 0.19 4 0.760 0.20 4 0.800 </code></pre> <p><hr></p> <h1>Pictures</h1> <p><strong><em>[Picture 1]</em></strong> : I've extended the search-space for $n$ to the range $\small 1 \ldots e^{1000} \approx 10^{434} $ and to check sanity to the range $\small 1 \ldots e^{2000} \approx 10^{868} $. For epsilons $\varepsilon$ from $0.001$ in $200$ steps up to $0.2$ I made the following graph:<br> <a href="https://i.stack.imgur.com/DAz5g.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/DAz5g.png" alt="picture"></a> </p> <p>Remarks: for the very small epsilon the increase of the search-space gives slightly higher results, which also illustrates, that for "larger" epsilon the cardinality of $\small M(\varepsilon)$ is finite</p> <p><hr> <strong><em>[Picture 2]</em></strong>: Indicates uniformity of the $\small f(n) =w_n = frac(h_n) \cdot n^1$ at the <em>n</em>, where $\small w_n $ has a local minimum (on request of @GerhardPaseman):<br> <a href="https://i.stack.imgur.com/18Az3.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/18Az3.png" alt="picture 2"></a> </p> <p><strong><em>[Picture 3]</em></strong>:It is also convenient to show a rescaling of the $\small f(n)$ so that we can immediately determine the cardinalities $\small |M(\varepsilon)|$ just by counting the number of dots $\small e(n)$ above $\small \varepsilon$. The derivation of the formula is </p> <p>$$ \small{ \begin{array}{lll} A(n) &amp;\lt &amp; {1\over n^1\cdot n^\varepsilon} &amp; \text{ from OP}\\ f(n) \cdot n^\varepsilon &amp;\lt &amp; 1\\ \ln(f(n)) + \varepsilon \cdot \ln(n) &amp;\lt &amp; 0 \\ {\ln(f(n)) \over \ln(n)} + \varepsilon &amp;\lt &amp; 0 \\ \varepsilon &amp; \lt&amp; -\ln_n(f(n)) \end{array} }$$</p> <p>and we simply count, how many dots in the picture show $\small e(n) = - \ln_n(f(n)) \gt \varepsilon$ </p> <p><a href="https://i.stack.imgur.com/QFKcS.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/QFKcS.png" alt="picture"></a> </p> <p><hr></p> <h1>Data for cardinalities-plot</h1> <pre><code>epsilon |M(eps,N)||M(eps,N)| N~e^1000 N~e^2000 ------------------------------- 0.0000 1000 2000 0.0010 633 862 0.0020 430 482 0.0030 315 330 0.0040 258 264 0.0050 204 205 0.0060 171 171 0.0070 151 151 0.0080 135 135 0.0090 120 120 0.0100 109 109 0.0110 100 100 0.0120 98 98 0.0130 89 89 0.0140 84 84 0.0150 79 79 0.0160 73 73 0.0170 69 69 0.0180 62 62 0.0190 57 57 0.0200 55 55 0.0210 50 50 0.0220 47 47 0.0230 44 44 0.0240 42 42 0.0250 41 41 0.0260 39 39 0.0270 39 39 0.0280 37 37 0.0290 37 37 0.0300 36 36 0.0310 36 36 0.0320 36 36 0.0330 36 36 0.0340 34 34 0.0350 34 34 0.0360 34 34 0.0370 33 33 0.0380 32 32 0.0390 31 31 0.0400 29 29 0.0410 29 29 0.0420 29 29 0.0430 26 26 0.0440 26 26 0.0450 25 25 0.0460 25 25 0.0470 24 24 0.0480 24 24 0.0490 23 23 0.0500 21 21 0.0510 21 21 0.0520 21 21 0.0530 20 20 0.0540 20 20 0.0550 20 20 0.0560 20 20 0.0570 19 19 0.0580 18 18 0.0590 18 18 0.0600 16 16 0.0610 16 16 0.0620 16 16 0.0630 16 16 0.0640 16 16 0.0650 16 16 0.0660 16 16 0.0670 16 16 0.0680 15 15 0.0690 14 14 0.0700 12 12 0.0710 12 12 0.0720 12 12 0.0730 12 12 0.0740 12 12 0.0750 12 12 0.0760 12 12 0.0770 12 12 0.0780 12 12 0.0790 12 12 0.0800 12 12 0.0810 11 11 0.0820 11 11 0.0830 11 11 0.0840 11 11 0.0850 11 11 0.0860 11 11 0.0870 11 11 0.0880 10 10 0.0890 10 10 0.0900 10 10 0.0910 10 10 0.0920 10 10 0.0930 10 10 0.0940 10 10 0.0950 10 10 0.0960 9 9 0.0970 9 9 0.0980 8 8 0.0990 8 8 0.1000 7 7 0.1010 7 7 0.1020 7 7 0.1030 7 7 0.1040 7 7 0.1050 7 7 0.1060 7 7 0.1070 7 7 0.1080 7 7 0.1090 6 6 0.1100 6 6 0.1110 6 6 0.1120 5 5 0.1130 5 5 0.1140 5 5 0.1150 4 4 0.1160 4 4 0.1170 4 4 0.1180 4 4 0.1190 4 4 0.1200 4 4 0.1210 4 4 0.1220 4 4 0.1230 4 4 0.1240 4 4 0.1250 4 4 0.1260 4 4 0.1270 4 4 0.1280 4 4 0.1290 4 4 0.1300 4 4 0.1310 4 4 0.1320 4 4 0.1330 4 4 0.1340 4 4 0.1350 4 4 0.1360 4 4 0.1370 4 4 0.1380 4 4 0.1390 4 4 0.1400 4 4 0.1410 4 4 0.1420 4 4 0.1430 4 4 0.1440 4 4 0.1450 4 4 0.1460 4 4 0.1470 4 4 0.1480 4 4 0.1490 4 4 0.1500 4 4 0.1510 4 4 0.1520 4 4 0.1530 4 4 0.1540 4 4 0.1550 4 4 0.1560 4 4 0.1570 4 4 0.1580 4 4 0.1590 4 4 0.1600 4 4 0.1610 4 4 0.1620 4 4 0.1630 4 4 0.1640 4 4 0.1650 4 4 0.1660 4 4 0.1670 4 4 0.1680 4 4 0.1690 4 4 0.1700 4 4 0.1710 4 4 0.1720 4 4 0.1730 4 4 0.1740 4 4 0.1750 4 4 0.1760 4 4 0.1770 4 4 0.1780 4 4 0.1790 4 4 0.1800 4 4 0.1810 4 4 0.1820 4 4 0.1830 4 4 0.1840 4 4 0.1850 4 4 0.1860 4 4 0.1870 4 4 0.1880 4 4 0.1890 4 4 0.1900 4 4 0.1910 4 4 0.1920 4 4 0.1930 4 4 0.1940 4 4 0.1950 4 4 0.1960 4 4 0.1970 4 4 0.1980 4 4 0.1990 4 4 </code></pre> Answer 4: <p>I am letting $A = H_n - \lfloor H_n \rfloor$ and computing it in a dumb way, and then computing $1/nA$ when $A \lt 1/n $. For $n=83$ I get $1/nA \approx 5.825$, which is the largest such value I find so far, and which means 83 is one of three values of $n$ which are allowed when $\epsilon$ is 0.3. A smarter way to compute the values of $n$ for which $1/nA$ is worth computing is to use $\log(n+1/2) + \gamma$ for just those $n$ for which the expression is very near an integer, for $H_n$ is within $1/24n^2$ of this value. $\log(1/nA )/\log(n)$ then serves as an upper bound on $\epsilon$.</p> <p>My calculations show $1 \lt 1/nA \lt 2$ much of the time, which means $\epsilon \lt 1/\log n$ for admissible $n$. I expect the conjecture to hold and when proved to reveal some of the nature of Euler's $\gamma$.</p> <p>Gerhard "Still Computing Oh So Slowly" Paseman, 2017.01.25.</p> Answer 5: <p><strong><em>This is a comment at the comments of Gerhard "Still Computing Oh So Slowly" Paseman's answer, giving just indexes n for more record-holders.</em></strong> </p> <p>Let $\small h_n$ denote the <em>n</em>'th harmonic number, $\small A_n=\{h_n\}$ its fractional part. The sequence of $A_n$ has a remarkable shape like a sawtooth-curve with increasing wavelength when <em>n</em> is increased and with sharp local minima - a shape which can be exploited when we seek for possible <em>n</em> to be included in <strong><em>M</em></strong>.<br> I use $\small f(n) = w_n = \{h_n\} \cdot n$ and $\small f(n,\varepsilon)=w_n \cdot n^\varepsilon $ with the OP's condition rewritten as $\small f(n,\varepsilon)&lt;1 $ as criterion for the inclusion of <em>n</em> into the set <strong><em>M</em></strong> . Of course the cardinality of <strong><em>M</em></strong> is limited by an upper bound $ \small n \le N$ with some <strong><em>N</em></strong> that can numerically be handled, so actually we should write explicitely $\small |M(\varepsilon,N)| $ instead of <strong><em>M</em></strong> only. I could manage to use $\small N \approx e^{2000}$ with the help of Pari/GP. </p> <p>The local minima of $\small A_n$ occur near $\small x_k=e^{k-\gamma}, k \in \mathbb N^+$ and the <em>n</em> to be tested is one of the next integers enclosing the $\small x_k$ - so this exact <em>n</em> must be empirically be determined. (Note that my <em>n</em> are the <em>n+1</em> used in the comments at the OP denoting the previous <em>high</em> value of <strong><em>A</em></strong> ) </p> <p>The harmonic numbers $\small h_n$ can be computed in Pari/GP using <code>h(n) = psi(1+n) + Euler</code> ; however, this seems to be limited to something like $ \small n \lt e^{600}$ and so I had to introduce the Euler-McLaurin-formula for the larger <em>n</em> and implemented the switch from one method to the other at <code>n = 1e50</code> </p> <p><hr> The following table shows the first <em>40</em> entries of my <em>2000</em>-row table containing $\small n \approx 10^{840}$ which was possible to compute using Pari/GP in just a couple of seconds. </p> <p>The basic option according to the focus in the OP's question is to test $\small f(n,\varepsilon) = w_n \cdot n^\varepsilon \lt 1$. For an example with $\small \varepsilon=0.1$ see the 6'th column and the 7'th column allowing to sum to the cardinality of <strong><em>M</em></strong> where we find $\small | M(0.1,2000) | = 7$ with $\small n \le N \approx e^{2000}$ </p> <p>A second, much nicer, option is to define the function $\small e(n) = -\log_n(w_n) $ and test $\small e(n) \gt \varepsilon$ whether to include this <em>n</em> or not. This is simply possible when looking into the <em>5</em>'th column and simply compare.<br> This latter method allows to compute the cardinalities of $\small M(\varepsilon)$ for arbitrary $\varepsilon$ really fast, for instance to create informative scatter- or lineplots, when first a list for $e(n)$ of length $\small N$ is created and then the successful comparisions with the intended $\varepsilon$ are summed to determine the cardinality. </p> <pre><code> n | h_n | A=frac(h_n) | w=A*n | e(n) | w*n^0.1|in M? -----------------+--------+-------------+----------+-----------+--------+--- 1 1.00000 0.00000 0.00000 1.000000 0.00000 1 4 2.08333 0.0833333 0.333333 0.792481 0.382899 1 11 3.01988 0.0198773 0.218651 0.634006 0.277901 1 31 4.02725 0.0272452 0.844601 0.0491822 1.19066 . 83 5.00207 0.00206827 0.171667 0.398793 0.267051 1 227 6.00437 0.00436671 0.991243 0.00162136 1.70523 . 616 7.00127 0.00127410 0.784844 0.0377178 1.49191 . 1674 8.00049 0.000485572 0.812848 0.0279149 1.70759 . 4550 9.00021 0.000208063 0.946686 0.00650460 2.19790 . 12367 10.0000 0.0000430083 0.531883 0.0670005 1.36472 . 33617 11.0000 0.0000177086 0.595311 0.0497632 1.68811 . 91380 12.0000 0.00000305167 0.278861 0.111798 0.873923 1 248397 13.0000 0.00000122948 0.305399 0.0954806 1.05775 . 675214 14.0000 1.36205E-7 0.919678 0.00623806 3.52030 . 1835421 15.0000 3.78268E-7 0.694281 0.0252988 2.93703 . 4989191 16.0000 9.54538E-7 0.476237 0.0481002 2.22652 . 13562027 17.0000 1.48499E-8 0.201395 0.0975770 1.04059 . 36865412 18.0000 3.71993E-9 0.137137 0.114033 0.783098 1 100210581 19.0000 9.73330E-9 0.975380 0.00135314 6.15552 . 272400600 20.0000 1.61744E-9 0.440592 0.0421997 3.07297 . 740461601 21.0000 4.01333E-9 0.297172 0.0594162 2.29065 . 2012783315 22.0000 1.38447E-10 0.278664 0.0596444 2.37389 . 5471312310 23.0000 1.97920E-11 0.108288 0.0991384 1.01951 . 14872568831 24.0000 2.27220E-11 0.337935 0.0463183 3.51618 . 40427833596 25.0000 6.07937E-12 0.245776 0.0574601 2.82623 . 109894245429 26.0000 7.60776E-12 0.836049 0.00704359 10.6250 . 298723530401 27.0000 1.82203E-12 0.544283 0.0230213 7.64454 . 812014744422 28.0000 5.52830E-13 0.448906 0.0292071 6.96806 . 2207284924203 29.0000 1.00870E-13 0.222650 0.0528504 3.81951 . 6000022499693 30.0000 2.16954E-14 0.130173 0.0692963 2.46795 . 16309752131262 31.0000 3.65111E-14 0.595487 0.0170391 12.4772 . 44334502845080 32.0000 1.81005E-15 0.080247 0.0802804 1.85827 . 120513673457548 33.0000 4.59281E-15 0.553496 0.0182434 14.1651 . 327590128640500 34.0000 2.31992E-15 0.759983 0.00821173 21.4950 . 890482293866031 35.0000 2.71425E-17 0.024169 0.108145 0.755506 1 2420581837980561 36.0000 2.55560E-16 0.618603 0.0135588 21.3700 . 6579823624480555 37.0000 1.20561E-17 0.079326 0.0695767 3.02860 . 17885814992891026 38.0000 4.26642E-18 0.076308 0.0687541 3.21976 . 48618685882356024 39.0000 7.23781E-19 0.035189 0.0871101 1.64093 . 132159290357566703 40.0000 2.02186E-18 0.267208 0.0334763 13.7708 . </code></pre> <p>The set $\small M(0.1,e^{40})$ (represented by <em>40</em> rows) is here the set of all $n$ where $\small f(n,0.1)=w_n \cdot n^{0.1} \lt 1$ . The number of such entries for $\small n=132159290357566703 \approx 10^{17}$ is here <em>7</em> . </p> <p>All numerical tests which I've done indicate that the cardinality of $\small M(\varepsilon \gt 0,\infty)$ is finite and roughly reciprocal to $\small \varepsilon$ and only if $\varepsilon=0$ is surely infinite. </p> <p><hr></p> <h1>Pari/GP tools</h1> <p>This is the Pari/GP-program which I used. </p> <p>Using the functions</p> <pre><code> h(n) = Euler+if(n&lt;1e50, psi(1+n),log(n)+1/2/n-1/12/n^2+1/120/n^4-1/252/n^6+1/240/n^8) A(n) = if(n==1,return(0)); frac(h(n)) w(n) = A(n)*n e(n) = if(n==1,return(1)); -log(w(n))/log(n) </code></pre> <p>The following needs only two steps to find the <em>n</em> with the next local minimum when called with index $\small k \in \mathbb N$ : </p> <pre><code> {find_n(k)=local(n1,n2,a1,a2); n1=floor(exp(k-Euler)); n2=n1+1; a1 = A(n1);a2 = A(n2); if(a1&lt;a2,return(n1),return(n2)); } \\ create that list one time, then evaluate cardinality for various eps \\ using that same list {makeList(listlen=40)= local(list,n,w1,logn); list=matrix(listlen,4); list[1,]=[Euler,0,0,0]; for(k=2,listlen, n=find_n(k); logn=log(n); list[k,]=[logn+Euler,n,w1=w(n),-log(w1)/logn]; ); return(list); } \\ compute cardinality with some eps cardM(eps,list) = sum(k=1,#list[,1],list[k,4]&gt;eps) \\ apply, note: for long lists we need high internal precision list = makeList(40) \\ the max N is here about e^40 print(cardM(0.1,list) ) </code></pre> <hr> <p><strong><em>[table 2]</em></strong>: This are sample data for the OP's plot of the cardinality of $\small M(\varepsilon)$ by the argument $\varepsilon$ where $\small N \approx \exp(250) \approx 10^{108}$ : </p> <pre><code> eps | c=#M | r=c*eps | c= card(M) for n&lt;= N approx 10^108 ------+------------------ 0.01 104 1.040 0.02 55 1.100 0.03 36 1.080 0.04 29 1.160 0.05 21 1.050 0.06 16 0.960 0.07 12 0.840 0.08 12 0.960 0.09 10 0.900 0.10 7 0.700 0.11 6 0.660 0.12 4 0.480 0.13 4 0.520 0.14 4 0.560 0.15 4 0.600 0.16 4 0.640 0.17 4 0.680 0.18 4 0.720 0.19 4 0.760 0.20 4 0.800 </code></pre> <p><hr></p> <h1>Pictures</h1> <p><strong><em>[Picture 1]</em></strong> : I've extended the search-space for $n$ to the range $\small 1 \ldots e^{1000} \approx 10^{434} $ and to check sanity to the range $\small 1 \ldots e^{2000} \approx 10^{868} $. For epsilons $\varepsilon$ from $0.001$ in $200$ steps up to $0.2$ I made the following graph:<br> <a href="https://i.stack.imgur.com/DAz5g.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/DAz5g.png" alt="picture"></a> </p> <p>Remarks: for the very small epsilon the increase of the search-space gives slightly higher results, which also illustrates, that for "larger" epsilon the cardinality of $\small M(\varepsilon)$ is finite</p> <p><hr> <strong><em>[Picture 2]</em></strong>: Indicates uniformity of the $\small f(n) =w_n = frac(h_n) \cdot n^1$ at the <em>n</em>, where $\small w_n $ has a local minimum (on request of @GerhardPaseman):<br> <a href="https://i.stack.imgur.com/18Az3.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/18Az3.png" alt="picture 2"></a> </p> <p><strong><em>[Picture 3]</em></strong>:It is also convenient to show a rescaling of the $\small f(n)$ so that we can immediately determine the cardinalities $\small |M(\varepsilon)|$ just by counting the number of dots $\small e(n)$ above $\small \varepsilon$. The derivation of the formula is </p> <p>$$ \small{ \begin{array}{lll} A(n) &amp;\lt &amp; {1\over n^1\cdot n^\varepsilon} &amp; \text{ from OP}\\ f(n) \cdot n^\varepsilon &amp;\lt &amp; 1\\ \ln(f(n)) + \varepsilon \cdot \ln(n) &amp;\lt &amp; 0 \\ {\ln(f(n)) \over \ln(n)} + \varepsilon &amp;\lt &amp; 0 \\ \varepsilon &amp; \lt&amp; -\ln_n(f(n)) \end{array} }$$</p> <p>and we simply count, how many dots in the picture show $\small e(n) = - \ln_n(f(n)) \gt \varepsilon$ </p> <p><a href="https://i.stack.imgur.com/QFKcS.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/QFKcS.png" alt="picture"></a> </p> <p><hr></p> <h1>Data for cardinalities-plot</h1> <pre><code>epsilon |M(eps,N)||M(eps,N)| N~e^1000 N~e^2000 ------------------------------- 0.0000 1000 2000 0.0010 633 862 0.0020 430 482 0.0030 315 330 0.0040 258 264 0.0050 204 205 0.0060 171 171 0.0070 151 151 0.0080 135 135 0.0090 120 120 0.0100 109 109 0.0110 100 100 0.0120 98 98 0.0130 89 89 0.0140 84 84 0.0150 79 79 0.0160 73 73 0.0170 69 69 0.0180 62 62 0.0190 57 57 0.0200 55 55 0.0210 50 50 0.0220 47 47 0.0230 44 44 0.0240 42 42 0.0250 41 41 0.0260 39 39 0.0270 39 39 0.0280 37 37 0.0290 37 37 0.0300 36 36 0.0310 36 36 0.0320 36 36 0.0330 36 36 0.0340 34 34 0.0350 34 34 0.0360 34 34 0.0370 33 33 0.0380 32 32 0.0390 31 31 0.0400 29 29 0.0410 29 29 0.0420 29 29 0.0430 26 26 0.0440 26 26 0.0450 25 25 0.0460 25 25 0.0470 24 24 0.0480 24 24 0.0490 23 23 0.0500 21 21 0.0510 21 21 0.0520 21 21 0.0530 20 20 0.0540 20 20 0.0550 20 20 0.0560 20 20 0.0570 19 19 0.0580 18 18 0.0590 18 18 0.0600 16 16 0.0610 16 16 0.0620 16 16 0.0630 16 16 0.0640 16 16 0.0650 16 16 0.0660 16 16 0.0670 16 16 0.0680 15 15 0.0690 14 14 0.0700 12 12 0.0710 12 12 0.0720 12 12 0.0730 12 12 0.0740 12 12 0.0750 12 12 0.0760 12 12 0.0770 12 12 0.0780 12 12 0.0790 12 12 0.0800 12 12 0.0810 11 11 0.0820 11 11 0.0830 11 11 0.0840 11 11 0.0850 11 11 0.0860 11 11 0.0870 11 11 0.0880 10 10 0.0890 10 10 0.0900 10 10 0.0910 10 10 0.0920 10 10 0.0930 10 10 0.0940 10 10 0.0950 10 10 0.0960 9 9 0.0970 9 9 0.0980 8 8 0.0990 8 8 0.1000 7 7 0.1010 7 7 0.1020 7 7 0.1030 7 7 0.1040 7 7 0.1050 7 7 0.1060 7 7 0.1070 7 7 0.1080 7 7 0.1090 6 6 0.1100 6 6 0.1110 6 6 0.1120 5 5 0.1130 5 5 0.1140 5 5 0.1150 4 4 0.1160 4 4 0.1170 4 4 0.1180 4 4 0.1190 4 4 0.1200 4 4 0.1210 4 4 0.1220 4 4 0.1230 4 4 0.1240 4 4 0.1250 4 4 0.1260 4 4 0.1270 4 4 0.1280 4 4 0.1290 4 4 0.1300 4 4 0.1310 4 4 0.1320 4 4 0.1330 4 4 0.1340 4 4 0.1350 4 4 0.1360 4 4 0.1370 4 4 0.1380 4 4 0.1390 4 4 0.1400 4 4 0.1410 4 4 0.1420 4 4 0.1430 4 4 0.1440 4 4 0.1450 4 4 0.1460 4 4 0.1470 4 4 0.1480 4 4 0.1490 4 4 0.1500 4 4 0.1510 4 4 0.1520 4 4 0.1530 4 4 0.1540 4 4 0.1550 4 4 0.1560 4 4 0.1570 4 4 0.1580 4 4 0.1590 4 4 0.1600 4 4 0.1610 4 4 0.1620 4 4 0.1630 4 4 0.1640 4 4 0.1650 4 4 0.1660 4 4 0.1670 4 4 0.1680 4 4 0.1690 4 4 0.1700 4 4 0.1710 4 4 0.1720 4 4 0.1730 4 4 0.1740 4 4 0.1750 4 4 0.1760 4 4 0.1770 4 4 0.1780 4 4 0.1790 4 4 0.1800 4 4 0.1810 4 4 0.1820 4 4 0.1830 4 4 0.1840 4 4 0.1850 4 4 0.1860 4 4 0.1870 4 4 0.1880 4 4 0.1890 4 4 0.1900 4 4 0.1910 4 4 0.1920 4 4 0.1930 4 4 0.1940 4 4 0.1950 4 4 0.1960 4 4 0.1970 4 4 0.1980 4 4 0.1990 4 4 </code></pre> Answer 6: <p>I am letting $A = H_n - \lfloor H_n \rfloor$ and computing it in a dumb way, and then computing $1/nA$ when $A \lt 1/n $. For $n=83$ I get $1/nA \approx 5.825$, which is the largest such value I find so far, and which means 83 is one of three values of $n$ which are allowed when $\epsilon$ is 0.3. A smarter way to compute the values of $n$ for which $1/nA$ is worth computing is to use $\log(n+1/2) + \gamma$ for just those $n$ for which the expression is very near an integer, for $H_n$ is within $1/24n^2$ of this value. $\log(1/nA )/\log(n)$ then serves as an upper bound on $\epsilon$.</p> <p>My calculations show $1 \lt 1/nA \lt 2$ much of the time, which means $\epsilon \lt 1/\log n$ for admissible $n$. I expect the conjecture to hold and when proved to reveal some of the nature of Euler's $\gamma$.</p> <p>Gerhard "Still Computing Oh So Slowly" Paseman, 2017.01.25.</p>
Relation between the homotopy classes of maps on a torus, and maps on a sphere
258,848
11
https://mathoverflow.net/questions/258848/relation-between-the-homotopy-classes-of-maps-on-a-torus-and-maps-on-a-sphere
<p>In modern condensed matter physics, one is often interested in the homotopy classes of mappings from a $d$-dimensional torus $$\mathbb{T}^d=\underbrace{S^1\times\ldots \times S^1}_d$$ (corresponding to the Brillouin zone of a $d$-dimensional system) to various topological spaces $Y$, for example the Grassmannians $$O(m+n)/O(m)\times O(n) \qquad \textrm{or} \qquad U(m+n)/U(m)\times U(n) $$ or some other symmetric spaces $-$ see for example Eq. (4) of <a href="http://aip.scitation.org/doi/pdf/10.1063/1.3149481" rel="noreferrer">this publication</a>. Such considerations lie at the hearth of the discussion of all topological phases of non-interacting particles, like the quantum Hall effect, Weyl semimetals or topological superconductors.</p> <p>Unfortunately, most texts on the topic are rather sloppy and without much explanation replace the torus $\mathbb{T}^d$ by a sphere $S^d$, thus instead considering the homotopy groups $\pi_d(Y)$.</p> <p>Hence the following questions: Is there any simple relation between the set of homotopy classes of mappings from $T^d$ to a general topological space $Y$ and the homotopy groups $\pi_d(Y)$? If the relation is in general too complex, under what conditions on $Y$ does it simplify? (For example, if $Y$ is a symmetric space?) Is the order of the two homotopy classes related?</p>
[ { "answer_id": 258854, "score": 13, "text": "<p>One case in which you can establish a simple relationship is when $Y$ is a loop space. Suppose that $Y\\simeq \\Omega Z=\\mbox{map}_*(S^1, Z)$. Then there is a bijection $[{\\mathbb T}^d, Y]_*\\cong [\\Sigma{\\mathbb T}^d, Z]_*$. On the other hand, there is an equivalence $$\\Sigma{\\mathbb T}^d \\simeq \\bigvee_{i\\ge 1} \\bigvee_{d\\choose i} S^{i+1}.$$</p>\n\n<p>This follows, using induction, from the equivalence $\\Sigma X\\times Y\\simeq \\Sigma X\\vee \\Sigma Y \\vee \\Sigma X\\wedge Y$ (Proposition 4I.1 in Hatcher's book). It follows that there is a bijection</p>\n\n<p>$$[{\\mathbb T}^d, Y]_*\\cong [\\bigvee_{1\\le i\\le d} \\bigvee_{d\\choose i} S^{i}, Y]_*\\cong \\prod_{i=1}^d\\prod_{d\\choose i} \\pi_i(Y).$$</p>\n\n<p>It is easy to extend this calculation to unpointed homotopy classes of maps, because $[{\\mathbb T}^d, Y]\\cong [{\\mathbb T}^d_+, Y]_*$ and there is an equivalence $\\Sigma {\\mathbb T}^d_+\\simeq \\Sigma {\\mathbb T}^d \\vee S^1$.</p>\n\n<p>Homogeneous spaces are usually not loop spaces, but for example topological groups are loop spaces, including the orthogonal and the unitary groups.</p>\n\n<p>By the way, this answer on math.SE has a very nice description of the set $[{\\mathbb T}^2, X]$ for a general space $X$. <a href=\"https://math.stackexchange.com/questions/36488/how-to-compute-homotopy-classes-of-maps-on-the-2-torus\">https://math.stackexchange.com/questions/36488/how-to-compute-homotopy-classes-of-maps-on-the-2-torus</a></p>\n" }, { "answer_id": 259021, "score": 1, "text": "<p>It would be interesting if the paper \n\"Homotopy Groups and Torus Homotopy Groups\", \nRalph H. Fox, \nAnnals of Mathematics, \nSecond Series, Vol. 49, No. 2 (Apr., 1948), pp. 471-510, \ncould be seen as relevant to aspects of physics! </p>\n\n<p>Actually I was told by Brian Griffiths that Fox was looking for higher order versions of the van Kampen theorem for the fundamental group but did not achieve that. That aim has been achieved by working on higher versions of <a href=\"http://groupoids.org.uk/gpdsweb.html\" rel=\"nofollow noreferrer\">groupoids</a>. </p>\n\n<p>I have also remembered the relevany paper \"Generalizations of Fox homotopy groups, Whitehead products, and Gottlieb groups\", \nM Golasinski, D Gonçalves, P Wong - Ukrainian Mathematical Journal, 2005 - Springer. </p>\n" }, { "answer_id": 258854, "score": 13, "text": "<p>One case in which you can establish a simple relationship is when $Y$ is a loop space. Suppose that $Y\\simeq \\Omega Z=\\mbox{map}_*(S^1, Z)$. Then there is a bijection $[{\\mathbb T}^d, Y]_*\\cong [\\Sigma{\\mathbb T}^d, Z]_*$. On the other hand, there is an equivalence $$\\Sigma{\\mathbb T}^d \\simeq \\bigvee_{i\\ge 1} \\bigvee_{d\\choose i} S^{i+1}.$$</p>\n\n<p>This follows, using induction, from the equivalence $\\Sigma X\\times Y\\simeq \\Sigma X\\vee \\Sigma Y \\vee \\Sigma X\\wedge Y$ (Proposition 4I.1 in Hatcher's book). It follows that there is a bijection</p>\n\n<p>$$[{\\mathbb T}^d, Y]_*\\cong [\\bigvee_{1\\le i\\le d} \\bigvee_{d\\choose i} S^{i}, Y]_*\\cong \\prod_{i=1}^d\\prod_{d\\choose i} \\pi_i(Y).$$</p>\n\n<p>It is easy to extend this calculation to unpointed homotopy classes of maps, because $[{\\mathbb T}^d, Y]\\cong [{\\mathbb T}^d_+, Y]_*$ and there is an equivalence $\\Sigma {\\mathbb T}^d_+\\simeq \\Sigma {\\mathbb T}^d \\vee S^1$.</p>\n\n<p>Homogeneous spaces are usually not loop spaces, but for example topological groups are loop spaces, including the orthogonal and the unitary groups.</p>\n\n<p>By the way, this answer on math.SE has a very nice description of the set $[{\\mathbb T}^2, X]$ for a general space $X$. <a href=\"https://math.stackexchange.com/questions/36488/how-to-compute-homotopy-classes-of-maps-on-the-2-torus\">https://math.stackexchange.com/questions/36488/how-to-compute-homotopy-classes-of-maps-on-the-2-torus</a></p>\n" }, { "answer_id": 259021, "score": 1, "text": "<p>It would be interesting if the paper \n\"Homotopy Groups and Torus Homotopy Groups\", \nRalph H. Fox, \nAnnals of Mathematics, \nSecond Series, Vol. 49, No. 2 (Apr., 1948), pp. 471-510, \ncould be seen as relevant to aspects of physics! </p>\n\n<p>Actually I was told by Brian Griffiths that Fox was looking for higher order versions of the van Kampen theorem for the fundamental group but did not achieve that. That aim has been achieved by working on higher versions of <a href=\"http://groupoids.org.uk/gpdsweb.html\" rel=\"nofollow noreferrer\">groupoids</a>. </p>\n\n<p>I have also remembered the relevany paper \"Generalizations of Fox homotopy groups, Whitehead products, and Gottlieb groups\", \nM Golasinski, D Gonçalves, P Wong - Ukrainian Mathematical Journal, 2005 - Springer. </p>\n" }, { "answer_id": 258854, "score": 13, "text": "<p>One case in which you can establish a simple relationship is when $Y$ is a loop space. Suppose that $Y\\simeq \\Omega Z=\\mbox{map}_*(S^1, Z)$. Then there is a bijection $[{\\mathbb T}^d, Y]_*\\cong [\\Sigma{\\mathbb T}^d, Z]_*$. On the other hand, there is an equivalence $$\\Sigma{\\mathbb T}^d \\simeq \\bigvee_{i\\ge 1} \\bigvee_{d\\choose i} S^{i+1}.$$</p>\n\n<p>This follows, using induction, from the equivalence $\\Sigma X\\times Y\\simeq \\Sigma X\\vee \\Sigma Y \\vee \\Sigma X\\wedge Y$ (Proposition 4I.1 in Hatcher's book). It follows that there is a bijection</p>\n\n<p>$$[{\\mathbb T}^d, Y]_*\\cong [\\bigvee_{1\\le i\\le d} \\bigvee_{d\\choose i} S^{i}, Y]_*\\cong \\prod_{i=1}^d\\prod_{d\\choose i} \\pi_i(Y).$$</p>\n\n<p>It is easy to extend this calculation to unpointed homotopy classes of maps, because $[{\\mathbb T}^d, Y]\\cong [{\\mathbb T}^d_+, Y]_*$ and there is an equivalence $\\Sigma {\\mathbb T}^d_+\\simeq \\Sigma {\\mathbb T}^d \\vee S^1$.</p>\n\n<p>Homogeneous spaces are usually not loop spaces, but for example topological groups are loop spaces, including the orthogonal and the unitary groups.</p>\n\n<p>By the way, this answer on math.SE has a very nice description of the set $[{\\mathbb T}^2, X]$ for a general space $X$. <a href=\"https://math.stackexchange.com/questions/36488/how-to-compute-homotopy-classes-of-maps-on-the-2-torus\">https://math.stackexchange.com/questions/36488/how-to-compute-homotopy-classes-of-maps-on-the-2-torus</a></p>\n" }, { "answer_id": 259021, "score": 1, "text": "<p>It would be interesting if the paper \n\"Homotopy Groups and Torus Homotopy Groups\", \nRalph H. Fox, \nAnnals of Mathematics, \nSecond Series, Vol. 49, No. 2 (Apr., 1948), pp. 471-510, \ncould be seen as relevant to aspects of physics! </p>\n\n<p>Actually I was told by Brian Griffiths that Fox was looking for higher order versions of the van Kampen theorem for the fundamental group but did not achieve that. That aim has been achieved by working on higher versions of <a href=\"http://groupoids.org.uk/gpdsweb.html\" rel=\"nofollow noreferrer\">groupoids</a>. </p>\n\n<p>I have also remembered the relevany paper \"Generalizations of Fox homotopy groups, Whitehead products, and Gottlieb groups\", \nM Golasinski, D Gonçalves, P Wong - Ukrainian Mathematical Journal, 2005 - Springer. </p>\n" } ]
Question: Relation between the homotopy classes of maps on a torus, and maps on a sphere Answer 1: <p>One case in which you can establish a simple relationship is when $Y$ is a loop space. Suppose that $Y\simeq \Omega Z=\mbox{map}_*(S^1, Z)$. Then there is a bijection $[{\mathbb T}^d, Y]_*\cong [\Sigma{\mathbb T}^d, Z]_*$. On the other hand, there is an equivalence $$\Sigma{\mathbb T}^d \simeq \bigvee_{i\ge 1} \bigvee_{d\choose i} S^{i+1}.$$</p> <p>This follows, using induction, from the equivalence $\Sigma X\times Y\simeq \Sigma X\vee \Sigma Y \vee \Sigma X\wedge Y$ (Proposition 4I.1 in Hatcher's book). It follows that there is a bijection</p> <p>$$[{\mathbb T}^d, Y]_*\cong [\bigvee_{1\le i\le d} \bigvee_{d\choose i} S^{i}, Y]_*\cong \prod_{i=1}^d\prod_{d\choose i} \pi_i(Y).$$</p> <p>It is easy to extend this calculation to unpointed homotopy classes of maps, because $[{\mathbb T}^d, Y]\cong [{\mathbb T}^d_+, Y]_*$ and there is an equivalence $\Sigma {\mathbb T}^d_+\simeq \Sigma {\mathbb T}^d \vee S^1$.</p> <p>Homogeneous spaces are usually not loop spaces, but for example topological groups are loop spaces, including the orthogonal and the unitary groups.</p> <p>By the way, this answer on math.SE has a very nice description of the set $[{\mathbb T}^2, X]$ for a general space $X$. <a href="https://math.stackexchange.com/questions/36488/how-to-compute-homotopy-classes-of-maps-on-the-2-torus">https://math.stackexchange.com/questions/36488/how-to-compute-homotopy-classes-of-maps-on-the-2-torus</a></p> Answer 2: <p>It would be interesting if the paper "Homotopy Groups and Torus Homotopy Groups", Ralph H. Fox, Annals of Mathematics, Second Series, Vol. 49, No. 2 (Apr., 1948), pp. 471-510, could be seen as relevant to aspects of physics! </p> <p>Actually I was told by Brian Griffiths that Fox was looking for higher order versions of the van Kampen theorem for the fundamental group but did not achieve that. That aim has been achieved by working on higher versions of <a href="http://groupoids.org.uk/gpdsweb.html" rel="nofollow noreferrer">groupoids</a>. </p> <p>I have also remembered the relevany paper "Generalizations of Fox homotopy groups, Whitehead products, and Gottlieb groups", M Golasinski, D Gonçalves, P Wong - Ukrainian Mathematical Journal, 2005 - Springer. </p> Answer 3: <p>One case in which you can establish a simple relationship is when $Y$ is a loop space. Suppose that $Y\simeq \Omega Z=\mbox{map}_*(S^1, Z)$. Then there is a bijection $[{\mathbb T}^d, Y]_*\cong [\Sigma{\mathbb T}^d, Z]_*$. On the other hand, there is an equivalence $$\Sigma{\mathbb T}^d \simeq \bigvee_{i\ge 1} \bigvee_{d\choose i} S^{i+1}.$$</p> <p>This follows, using induction, from the equivalence $\Sigma X\times Y\simeq \Sigma X\vee \Sigma Y \vee \Sigma X\wedge Y$ (Proposition 4I.1 in Hatcher's book). It follows that there is a bijection</p> <p>$$[{\mathbb T}^d, Y]_*\cong [\bigvee_{1\le i\le d} \bigvee_{d\choose i} S^{i}, Y]_*\cong \prod_{i=1}^d\prod_{d\choose i} \pi_i(Y).$$</p> <p>It is easy to extend this calculation to unpointed homotopy classes of maps, because $[{\mathbb T}^d, Y]\cong [{\mathbb T}^d_+, Y]_*$ and there is an equivalence $\Sigma {\mathbb T}^d_+\simeq \Sigma {\mathbb T}^d \vee S^1$.</p> <p>Homogeneous spaces are usually not loop spaces, but for example topological groups are loop spaces, including the orthogonal and the unitary groups.</p> <p>By the way, this answer on math.SE has a very nice description of the set $[{\mathbb T}^2, X]$ for a general space $X$. <a href="https://math.stackexchange.com/questions/36488/how-to-compute-homotopy-classes-of-maps-on-the-2-torus">https://math.stackexchange.com/questions/36488/how-to-compute-homotopy-classes-of-maps-on-the-2-torus</a></p> Answer 4: <p>It would be interesting if the paper "Homotopy Groups and Torus Homotopy Groups", Ralph H. Fox, Annals of Mathematics, Second Series, Vol. 49, No. 2 (Apr., 1948), pp. 471-510, could be seen as relevant to aspects of physics! </p> <p>Actually I was told by Brian Griffiths that Fox was looking for higher order versions of the van Kampen theorem for the fundamental group but did not achieve that. That aim has been achieved by working on higher versions of <a href="http://groupoids.org.uk/gpdsweb.html" rel="nofollow noreferrer">groupoids</a>. </p> <p>I have also remembered the relevany paper "Generalizations of Fox homotopy groups, Whitehead products, and Gottlieb groups", M Golasinski, D Gonçalves, P Wong - Ukrainian Mathematical Journal, 2005 - Springer. </p> Answer 5: <p>One case in which you can establish a simple relationship is when $Y$ is a loop space. Suppose that $Y\simeq \Omega Z=\mbox{map}_*(S^1, Z)$. Then there is a bijection $[{\mathbb T}^d, Y]_*\cong [\Sigma{\mathbb T}^d, Z]_*$. On the other hand, there is an equivalence $$\Sigma{\mathbb T}^d \simeq \bigvee_{i\ge 1} \bigvee_{d\choose i} S^{i+1}.$$</p> <p>This follows, using induction, from the equivalence $\Sigma X\times Y\simeq \Sigma X\vee \Sigma Y \vee \Sigma X\wedge Y$ (Proposition 4I.1 in Hatcher's book). It follows that there is a bijection</p> <p>$$[{\mathbb T}^d, Y]_*\cong [\bigvee_{1\le i\le d} \bigvee_{d\choose i} S^{i}, Y]_*\cong \prod_{i=1}^d\prod_{d\choose i} \pi_i(Y).$$</p> <p>It is easy to extend this calculation to unpointed homotopy classes of maps, because $[{\mathbb T}^d, Y]\cong [{\mathbb T}^d_+, Y]_*$ and there is an equivalence $\Sigma {\mathbb T}^d_+\simeq \Sigma {\mathbb T}^d \vee S^1$.</p> <p>Homogeneous spaces are usually not loop spaces, but for example topological groups are loop spaces, including the orthogonal and the unitary groups.</p> <p>By the way, this answer on math.SE has a very nice description of the set $[{\mathbb T}^2, X]$ for a general space $X$. <a href="https://math.stackexchange.com/questions/36488/how-to-compute-homotopy-classes-of-maps-on-the-2-torus">https://math.stackexchange.com/questions/36488/how-to-compute-homotopy-classes-of-maps-on-the-2-torus</a></p> Answer 6: <p>It would be interesting if the paper "Homotopy Groups and Torus Homotopy Groups", Ralph H. Fox, Annals of Mathematics, Second Series, Vol. 49, No. 2 (Apr., 1948), pp. 471-510, could be seen as relevant to aspects of physics! </p> <p>Actually I was told by Brian Griffiths that Fox was looking for higher order versions of the van Kampen theorem for the fundamental group but did not achieve that. That aim has been achieved by working on higher versions of <a href="http://groupoids.org.uk/gpdsweb.html" rel="nofollow noreferrer">groupoids</a>. </p> <p>I have also remembered the relevany paper "Generalizations of Fox homotopy groups, Whitehead products, and Gottlieb groups", M Golasinski, D Gonçalves, P Wong - Ukrainian Mathematical Journal, 2005 - Springer. </p>
A neat evaluation of an infinite matrix?
259,104
11
https://mathoverflow.net/questions/259104/a-neat-evaluation-of-an-infinite-matrix
<p>Let $M_n$ be an $n\times n$ matrix defined as $$M_n =\left[\frac{2i+1}{2(i+j+1)}\binom{i-1/2}i\binom{j-1/2}jx^{i+j+1}\right]_{i,j=0}^n.$$ With $I_n$ the identity matrix, consider $A_n:=I_n-M_n^2$. When I computed $\det A_n$, it looks rather "ugly". However, its infinite dimensional counterpart $\det A_{\infty}$ seems to reach a neat evaluation. To avoid issues on what it means by "determinant of infinite matrix", I simply work with the following convention: $\det(A_{\infty})=\lim_{n\rightarrow\infty}\det(A_n)$. So,</p> <blockquote> <p><strong>Question.</strong> Is this determinantal evaluation true? $$\det(A_{\infty})=\sqrt[4]{1-x^2}.$$</p> </blockquote> <p><strong>NOTE 1.</strong> The fractional values $\binom{i-1/2}i$ are (as usual) computed via Euler's Gamma function, $\Gamma(z)$.</p> <p><strong>NOTE 2.</strong> If it helps, we make two observations: (a) both $\det A_n$ and $\sqrt[4]{1-x^2}$ are functions of $y:=x^2$; (b) as functions of $y$, the taylor series for $\det A_n$ and $\sqrt[4]{1-y}$ agree up to degree $n$.</p>
[ { "answer_id": 381574, "score": 7, "text": "<p>The claimed expression for the determinant indeed holds. In fact, one may explicitly determine the eigenvalue decomposition of the infinite matrix <span class=\"math-container\">$M_\\infty$</span> viewed as a self-adjoint operator on a Hilbert space. Its eigenvalues are\n<span class=\"math-container\">$$ \\left(\\frac{2}{q_x^{p+\\frac12}+q_x^{-p-\\frac12}}\\right)_{p=0}^\\infty,\\qquad q_k = e^{-\\pi\\frac{K\\left(\\sqrt{1-x^2}\\right)}{K(x)}}, \\tag{1}\n$$</span> where <span class=\"math-container\">$q_k$</span> is the elliptic nome of modulus <span class=\"math-container\">$x$</span> and <span class=\"math-container\">$K(x)$</span> the complete elliptic integral of the first kind.\nOnce this is known one may compute\n<span class=\"math-container\">$$\n\\begin{aligned}\n\\det(I - M_\\infty^2) &amp;= \\prod_{p=0}^\\infty\\left(1-\\left(\\frac{2}{q_x^{p+\\frac12}+q_x^{-p-\\frac12}}\\right)^2\\right) \\\\\n&amp;=\\prod_{p=0}^\\infty\\left(\\frac{1-q_x^{2p+1}}{1+q_x^{2p+1}}\\right)^2 = \\frac{\\theta_4(0,q_x)}{\\theta_3(0,q_x)}=\\sqrt[4]{1-x^2},\n\\end{aligned}\n$$</span>\nwhere the last two equalities follow from standard properties of the <a href=\"https://mathworld.wolfram.com/JacobiThetaFunctions.html\" rel=\"nofollow noreferrer\">Jacobi theta functions <span class=\"math-container\">$\\theta_i$</span></a>.</p>\n<p>To see how the eigenvalue decomposition can be obtained, it is convenient to look at the Dirichlet space <span class=\"math-container\">$\\mathcal{D}$</span> of complex analytic functions <span class=\"math-container\">$f$</span> on the open unit disk that vanish at <span class=\"math-container\">$0$</span> and that have finite norm with respect to the Dirichlet inner product\n<span class=\"math-container\">$$\\langle f,g\\rangle_{\\mathcal{D}} = \\frac{1}{\\pi} \\int_{|z|&lt;1} \\overline{f'(z)}g'(z) \\mathrm{d}^2z = \\sum_{n=1}^\\infty n\\,\\overline{[z^n]f(z)}\\,[z^n]g(z).$$</span>\nA basis is given by the monomials <span class=\"math-container\">$(e_n:=z^n)_{n\\geq 1}$</span>, which satisfy <span class=\"math-container\">$\\langle e_n,e_m\\rangle_{\\mathcal{D}}=n\\, \\delta_{n,m}$</span>. Let us also introduce the bounded linear operator on <span class=\"math-container\">$\\mathcal{D}$</span> given by\n<span class=\"math-container\">$$\\mathbf{R}_x f = f \\circ \\psi_x, \\qquad \\psi_x(z) = \\frac{1-\\sqrt{1-x z^2}}{\\sqrt{x}\\,z}.$$</span>\nBy power series expansion one may check that\n<span class=\"math-container\">$$ \\mathbf{R}_x e_p = \\sum_{\\ell=p}^\\infty \\left(\\frac{x}{4}\\right)^{\\ell/2} \\frac{p}{\\ell}\\binom{\\ell}{(\\ell+p)/2}\\, 1_{\\{\\ell+p\\text{ even}\\}} e_\\ell$$</span>\nand the adjoint with respect to <span class=\"math-container\">$\\langle \\cdot,\\cdot\\rangle_{\\mathcal{D}}$</span> is determined by\n<span class=\"math-container\">$$ \\mathbf{R}_x^\\dagger e_p = \\sum_{\\ell=1}^p \\left(\\frac{x}{4}\\right)^{p/2} \\binom{p}{(\\ell+p)/2}\\, 1_{\\{\\ell+p\\text{ even}\\}} e_\\ell.$$</span>\nIn particular, <span class=\"math-container\">$\\mathbf{R}_x\\mathbf{R}_x^\\dagger$</span> is a self-adjoint operator that preserves the even and odd functions in <span class=\"math-container\">$\\mathcal{D}$</span>.\nUp to a factor of <span class=\"math-container\">$2$</span> its matrix elements on the odd monomials are precisely <span class=\"math-container\">$[M_\\infty]_{i,j}$</span>:\n<span class=\"math-container\">$$\n\\begin{aligned}\n2\\mathbf{R}_x\\mathbf{R}_x^\\dagger e_{2i+1} &amp;= \\sum_{j=0}^\\infty \\sum_{\\ell=0}^{\\min(i,j)} 2\\left(\\frac{x}{4}\\right)^{i+j+1} \\binom{2i+1}{i+\\ell+1}\\frac{2\\ell+1}{2j+1} \\binom{2j+1}{j+\\ell+1} e_{2j+1}\\\\\n&amp;= \\sum_{j=0}^\\infty\\frac{2i+1}{2(i+j+1)}\\binom{i-1/2}i\\binom{j-1/2}jx^{i+j+1}e_{2j+1}\\\\\n&amp;= \\sum_{j=0}^\\infty [M_\\infty]_{i,j} e_{2j+1}.\n\\end{aligned}\n$$</span>\nIn <em>T. Budd</em>, <a href=\"http://dx.doi.org/10.1016/j.jcta.2019.105191\" rel=\"nofollow noreferrer\"><strong>Winding of simple walks on the square lattice</strong></a>, <a href=\"https://arxiv.org/abs/1709.04042\" rel=\"nofollow noreferrer\">arXiv:1709.04042</a>, Section 2.1, I examined a closely related operator <span class=\"math-container\">$\\mathbf{J}_x = \\mathbf{R}_x^\\dagger\\mathbf{R}_x$</span>, whose matrix elements count diagonal walks on <span class=\"math-container\">$\\mathbb{Z}^2$</span> starting on the positive <span class=\"math-container\">$x$</span>-axis and hitting the <span class=\"math-container\">$y$</span>-axis at prescribed height. By a convenient elliptic parametrization of the disc one can determine all eigenvectors of <span class=\"math-container\">$\\mathbf{J}_x$</span>. Since <span class=\"math-container\">$\\mathbf{R}_x$</span> is injective, <span class=\"math-container\">$\\mathbf{R}_x\\mathbf{R}_x^\\dagger$</span> has the same eigenvalue decomposition (after applying <span class=\"math-container\">$\\mathbf{R}_x$</span> to the eigenvectors of <span class=\"math-container\">$\\mathbf{J}_x$</span>) given by Proposition 9, of which the eigenvectors with odd label <span class=\"math-container\">$m=1,3,5,\\ldots$</span>\nspan the odd functions in <span class=\"math-container\">$\\mathcal{D}$</span>. The corresponding eigenvalues are precisely the ones given above in (1).</p>\n<p><strong>Added (22 Jan):</strong> One may also check the formula for <span class=\"math-container\">$\\det(I\\pm M_\\infty)$</span> proposed in the <a href=\"https://mathoverflow.net/a/325953/47484\">post by Hucht</a>. Using the <a href=\"https://dlmf.nist.gov/20.4\" rel=\"nofollow noreferrer\">theta function product formulas</a> we find\n<span class=\"math-container\">$$\n\\det(I-M_\\infty) = \\prod_{p=0}^\\infty \\frac{(1-q_x^{p+\\frac12})^2}{1+q^{2p+1}} = \\left(\\frac{\\theta_2(0,\\sqrt{q_x})}{\\theta_1'(0,\\sqrt{q_x})}\\right)^{\\frac14} \\frac{\\theta_4(0,\\sqrt{q_x})}{\\theta_3(0,q_x)^{\\frac12}}.\n$$</span>\nWith the help of <span class=\"math-container\">$\\theta_1'(0,\\sqrt{q_x}) = \\theta_2(0,\\sqrt{q_x})\\theta_3(0,\\sqrt{q_x})\\theta_4(0,\\sqrt{q_x})$</span> and\n<span class=\"math-container\">$$\n\\theta_3(0,\\sqrt{q_x})^2 = (1+x)\\theta_3(0,q_x)^2,\\quad \\theta_4(0,\\sqrt{q_x})^2 = (1-x)\\theta_3(0,q_x)^2\n$$</span>\nthis yields\n<span class=\"math-container\">$$\n\\det(I- M_\\infty) = \\left(\\frac{(1-x)^3}{1+x}\\right)^{1/8}.\n$$</span>\nSimilarly\n<span class=\"math-container\">$$\n\\det(I-M_\\infty) = \\prod_{p=0}^\\infty \\frac{(1+q_x^{p+\\frac12})^2}{1+q^{2p+1}} = \\left(\\frac{\\theta_2(0,\\sqrt{q_x})}{\\theta_1'(0,\\sqrt{q_x})}\\right)^{\\frac14} \\frac{\\theta_3(0,\\sqrt{q_x})}{\\theta_3(0,q_x)^{\\frac12}}= \\left(\\frac{(1+x)^3}{1-x}\\right)^{1/8}.\n$$</span></p>\n" }, { "answer_id": 325953, "score": 2, "text": "<p>Based on your comment to my posting \"<a href=\"https://mathoverflow.net/questions/325886\">Conjecture for a certain Cauchy-type determinant</a>\", I found that in addition to the formula above (I write <span class=\"math-container\">$q$</span> instead of <span class=\"math-container\">$x$</span>, as all this is related to <span class=\"math-container\">$q$</span>-series, <span class=\"math-container\">$q$</span>-products, and Jacobi elliptic functions)\n<span class=\"math-container\">$$\n\\det[\\mathbf I_\\infty - \\mathbf M_\\infty(q) \\, \\mathbf M_\\infty(q)]\n=(1-q^2)^{1/4} \\qquad(1)\n$$</span>\nanother identity holds, which is closer to my case, namely\n<span class=\"math-container\">$$\n\\det[\\mathbf I_\\infty + \\mathbf M_\\infty^T(q) \\,\\mathbf M_\\infty(q)]\n=(1-q^2)^{-1/4}. \\qquad(2)\n$$</span>\nPlaying around a little further using my favorite tool <em>Mathematica</em>, I found the two \"roots\" of (1),\n<span class=\"math-container\">$$\n\\det[\\mathbf I_\\infty \\pm \\mathbf M_\\infty(q)] = \n\\left[\\frac{(1 \\pm q)^{3}}{1 \\mp q}\\right]^{1/8}.\n$$</span>\nI think that the resulting identity \n<span class=\"math-container\">$$\n\\det[\\mathbf I_\\infty + \\mathbf M_\\infty(q)] = \n\\det[\\mathbf I_\\infty - \\mathbf M_\\infty(-q)]\n$$</span>\ncan be proven quite easily by analyzing the series expansions.</p>\n\n<p>As <span class=\"math-container\">$\\mathbf M_\\infty$</span> and <span class=\"math-container\">$\\mathbf M_\\infty^T$</span> do not commute, (2) cannot easily be factored. \nHowever, maybe this helps a bit.</p>\n\n<p>Another observation is, that my problem <a href=\"https://mathoverflow.net/questions/325886\">325886</a> is an infinite product of this one, as\n<span class=\"math-container\">$$\n\\det[\\mathbf I_\\infty + \\mathbf X_\\infty^T(q) \\,\\mathbf X_\\infty(q)]\n= \\prod_{k=1}^{\\infty} \n\\det[\\mathbf I_\\infty + \\mathbf M_\\infty^T(q^k) \\,\\mathbf M_\\infty(q^k)].\n$$</span>\nSo they are definitely related. Obviously, all these expressions are matrix versions of <span class=\"math-container\">$q$</span>-product formulas. For further observations, see my updated posting <a href=\"https://mathoverflow.net/questions/325886\">325886</a>.</p>\n" }, { "answer_id": 381574, "score": 7, "text": "<p>The claimed expression for the determinant indeed holds. In fact, one may explicitly determine the eigenvalue decomposition of the infinite matrix <span class=\"math-container\">$M_\\infty$</span> viewed as a self-adjoint operator on a Hilbert space. Its eigenvalues are\n<span class=\"math-container\">$$ \\left(\\frac{2}{q_x^{p+\\frac12}+q_x^{-p-\\frac12}}\\right)_{p=0}^\\infty,\\qquad q_k = e^{-\\pi\\frac{K\\left(\\sqrt{1-x^2}\\right)}{K(x)}}, \\tag{1}\n$$</span> where <span class=\"math-container\">$q_k$</span> is the elliptic nome of modulus <span class=\"math-container\">$x$</span> and <span class=\"math-container\">$K(x)$</span> the complete elliptic integral of the first kind.\nOnce this is known one may compute\n<span class=\"math-container\">$$\n\\begin{aligned}\n\\det(I - M_\\infty^2) &amp;= \\prod_{p=0}^\\infty\\left(1-\\left(\\frac{2}{q_x^{p+\\frac12}+q_x^{-p-\\frac12}}\\right)^2\\right) \\\\\n&amp;=\\prod_{p=0}^\\infty\\left(\\frac{1-q_x^{2p+1}}{1+q_x^{2p+1}}\\right)^2 = \\frac{\\theta_4(0,q_x)}{\\theta_3(0,q_x)}=\\sqrt[4]{1-x^2},\n\\end{aligned}\n$$</span>\nwhere the last two equalities follow from standard properties of the <a href=\"https://mathworld.wolfram.com/JacobiThetaFunctions.html\" rel=\"nofollow noreferrer\">Jacobi theta functions <span class=\"math-container\">$\\theta_i$</span></a>.</p>\n<p>To see how the eigenvalue decomposition can be obtained, it is convenient to look at the Dirichlet space <span class=\"math-container\">$\\mathcal{D}$</span> of complex analytic functions <span class=\"math-container\">$f$</span> on the open unit disk that vanish at <span class=\"math-container\">$0$</span> and that have finite norm with respect to the Dirichlet inner product\n<span class=\"math-container\">$$\\langle f,g\\rangle_{\\mathcal{D}} = \\frac{1}{\\pi} \\int_{|z|&lt;1} \\overline{f'(z)}g'(z) \\mathrm{d}^2z = \\sum_{n=1}^\\infty n\\,\\overline{[z^n]f(z)}\\,[z^n]g(z).$$</span>\nA basis is given by the monomials <span class=\"math-container\">$(e_n:=z^n)_{n\\geq 1}$</span>, which satisfy <span class=\"math-container\">$\\langle e_n,e_m\\rangle_{\\mathcal{D}}=n\\, \\delta_{n,m}$</span>. Let us also introduce the bounded linear operator on <span class=\"math-container\">$\\mathcal{D}$</span> given by\n<span class=\"math-container\">$$\\mathbf{R}_x f = f \\circ \\psi_x, \\qquad \\psi_x(z) = \\frac{1-\\sqrt{1-x z^2}}{\\sqrt{x}\\,z}.$$</span>\nBy power series expansion one may check that\n<span class=\"math-container\">$$ \\mathbf{R}_x e_p = \\sum_{\\ell=p}^\\infty \\left(\\frac{x}{4}\\right)^{\\ell/2} \\frac{p}{\\ell}\\binom{\\ell}{(\\ell+p)/2}\\, 1_{\\{\\ell+p\\text{ even}\\}} e_\\ell$$</span>\nand the adjoint with respect to <span class=\"math-container\">$\\langle \\cdot,\\cdot\\rangle_{\\mathcal{D}}$</span> is determined by\n<span class=\"math-container\">$$ \\mathbf{R}_x^\\dagger e_p = \\sum_{\\ell=1}^p \\left(\\frac{x}{4}\\right)^{p/2} \\binom{p}{(\\ell+p)/2}\\, 1_{\\{\\ell+p\\text{ even}\\}} e_\\ell.$$</span>\nIn particular, <span class=\"math-container\">$\\mathbf{R}_x\\mathbf{R}_x^\\dagger$</span> is a self-adjoint operator that preserves the even and odd functions in <span class=\"math-container\">$\\mathcal{D}$</span>.\nUp to a factor of <span class=\"math-container\">$2$</span> its matrix elements on the odd monomials are precisely <span class=\"math-container\">$[M_\\infty]_{i,j}$</span>:\n<span class=\"math-container\">$$\n\\begin{aligned}\n2\\mathbf{R}_x\\mathbf{R}_x^\\dagger e_{2i+1} &amp;= \\sum_{j=0}^\\infty \\sum_{\\ell=0}^{\\min(i,j)} 2\\left(\\frac{x}{4}\\right)^{i+j+1} \\binom{2i+1}{i+\\ell+1}\\frac{2\\ell+1}{2j+1} \\binom{2j+1}{j+\\ell+1} e_{2j+1}\\\\\n&amp;= \\sum_{j=0}^\\infty\\frac{2i+1}{2(i+j+1)}\\binom{i-1/2}i\\binom{j-1/2}jx^{i+j+1}e_{2j+1}\\\\\n&amp;= \\sum_{j=0}^\\infty [M_\\infty]_{i,j} e_{2j+1}.\n\\end{aligned}\n$$</span>\nIn <em>T. Budd</em>, <a href=\"http://dx.doi.org/10.1016/j.jcta.2019.105191\" rel=\"nofollow noreferrer\"><strong>Winding of simple walks on the square lattice</strong></a>, <a href=\"https://arxiv.org/abs/1709.04042\" rel=\"nofollow noreferrer\">arXiv:1709.04042</a>, Section 2.1, I examined a closely related operator <span class=\"math-container\">$\\mathbf{J}_x = \\mathbf{R}_x^\\dagger\\mathbf{R}_x$</span>, whose matrix elements count diagonal walks on <span class=\"math-container\">$\\mathbb{Z}^2$</span> starting on the positive <span class=\"math-container\">$x$</span>-axis and hitting the <span class=\"math-container\">$y$</span>-axis at prescribed height. By a convenient elliptic parametrization of the disc one can determine all eigenvectors of <span class=\"math-container\">$\\mathbf{J}_x$</span>. Since <span class=\"math-container\">$\\mathbf{R}_x$</span> is injective, <span class=\"math-container\">$\\mathbf{R}_x\\mathbf{R}_x^\\dagger$</span> has the same eigenvalue decomposition (after applying <span class=\"math-container\">$\\mathbf{R}_x$</span> to the eigenvectors of <span class=\"math-container\">$\\mathbf{J}_x$</span>) given by Proposition 9, of which the eigenvectors with odd label <span class=\"math-container\">$m=1,3,5,\\ldots$</span>\nspan the odd functions in <span class=\"math-container\">$\\mathcal{D}$</span>. The corresponding eigenvalues are precisely the ones given above in (1).</p>\n<p><strong>Added (22 Jan):</strong> One may also check the formula for <span class=\"math-container\">$\\det(I\\pm M_\\infty)$</span> proposed in the <a href=\"https://mathoverflow.net/a/325953/47484\">post by Hucht</a>. Using the <a href=\"https://dlmf.nist.gov/20.4\" rel=\"nofollow noreferrer\">theta function product formulas</a> we find\n<span class=\"math-container\">$$\n\\det(I-M_\\infty) = \\prod_{p=0}^\\infty \\frac{(1-q_x^{p+\\frac12})^2}{1+q^{2p+1}} = \\left(\\frac{\\theta_2(0,\\sqrt{q_x})}{\\theta_1'(0,\\sqrt{q_x})}\\right)^{\\frac14} \\frac{\\theta_4(0,\\sqrt{q_x})}{\\theta_3(0,q_x)^{\\frac12}}.\n$$</span>\nWith the help of <span class=\"math-container\">$\\theta_1'(0,\\sqrt{q_x}) = \\theta_2(0,\\sqrt{q_x})\\theta_3(0,\\sqrt{q_x})\\theta_4(0,\\sqrt{q_x})$</span> and\n<span class=\"math-container\">$$\n\\theta_3(0,\\sqrt{q_x})^2 = (1+x)\\theta_3(0,q_x)^2,\\quad \\theta_4(0,\\sqrt{q_x})^2 = (1-x)\\theta_3(0,q_x)^2\n$$</span>\nthis yields\n<span class=\"math-container\">$$\n\\det(I- M_\\infty) = \\left(\\frac{(1-x)^3}{1+x}\\right)^{1/8}.\n$$</span>\nSimilarly\n<span class=\"math-container\">$$\n\\det(I-M_\\infty) = \\prod_{p=0}^\\infty \\frac{(1+q_x^{p+\\frac12})^2}{1+q^{2p+1}} = \\left(\\frac{\\theta_2(0,\\sqrt{q_x})}{\\theta_1'(0,\\sqrt{q_x})}\\right)^{\\frac14} \\frac{\\theta_3(0,\\sqrt{q_x})}{\\theta_3(0,q_x)^{\\frac12}}= \\left(\\frac{(1+x)^3}{1-x}\\right)^{1/8}.\n$$</span></p>\n" }, { "answer_id": 325953, "score": 2, "text": "<p>Based on your comment to my posting \"<a href=\"https://mathoverflow.net/questions/325886\">Conjecture for a certain Cauchy-type determinant</a>\", I found that in addition to the formula above (I write <span class=\"math-container\">$q$</span> instead of <span class=\"math-container\">$x$</span>, as all this is related to <span class=\"math-container\">$q$</span>-series, <span class=\"math-container\">$q$</span>-products, and Jacobi elliptic functions)\n<span class=\"math-container\">$$\n\\det[\\mathbf I_\\infty - \\mathbf M_\\infty(q) \\, \\mathbf M_\\infty(q)]\n=(1-q^2)^{1/4} \\qquad(1)\n$$</span>\nanother identity holds, which is closer to my case, namely\n<span class=\"math-container\">$$\n\\det[\\mathbf I_\\infty + \\mathbf M_\\infty^T(q) \\,\\mathbf M_\\infty(q)]\n=(1-q^2)^{-1/4}. \\qquad(2)\n$$</span>\nPlaying around a little further using my favorite tool <em>Mathematica</em>, I found the two \"roots\" of (1),\n<span class=\"math-container\">$$\n\\det[\\mathbf I_\\infty \\pm \\mathbf M_\\infty(q)] = \n\\left[\\frac{(1 \\pm q)^{3}}{1 \\mp q}\\right]^{1/8}.\n$$</span>\nI think that the resulting identity \n<span class=\"math-container\">$$\n\\det[\\mathbf I_\\infty + \\mathbf M_\\infty(q)] = \n\\det[\\mathbf I_\\infty - \\mathbf M_\\infty(-q)]\n$$</span>\ncan be proven quite easily by analyzing the series expansions.</p>\n\n<p>As <span class=\"math-container\">$\\mathbf M_\\infty$</span> and <span class=\"math-container\">$\\mathbf M_\\infty^T$</span> do not commute, (2) cannot easily be factored. \nHowever, maybe this helps a bit.</p>\n\n<p>Another observation is, that my problem <a href=\"https://mathoverflow.net/questions/325886\">325886</a> is an infinite product of this one, as\n<span class=\"math-container\">$$\n\\det[\\mathbf I_\\infty + \\mathbf X_\\infty^T(q) \\,\\mathbf X_\\infty(q)]\n= \\prod_{k=1}^{\\infty} \n\\det[\\mathbf I_\\infty + \\mathbf M_\\infty^T(q^k) \\,\\mathbf M_\\infty(q^k)].\n$$</span>\nSo they are definitely related. Obviously, all these expressions are matrix versions of <span class=\"math-container\">$q$</span>-product formulas. For further observations, see my updated posting <a href=\"https://mathoverflow.net/questions/325886\">325886</a>.</p>\n" }, { "answer_id": 381574, "score": 7, "text": "<p>The claimed expression for the determinant indeed holds. In fact, one may explicitly determine the eigenvalue decomposition of the infinite matrix <span class=\"math-container\">$M_\\infty$</span> viewed as a self-adjoint operator on a Hilbert space. Its eigenvalues are\n<span class=\"math-container\">$$ \\left(\\frac{2}{q_x^{p+\\frac12}+q_x^{-p-\\frac12}}\\right)_{p=0}^\\infty,\\qquad q_k = e^{-\\pi\\frac{K\\left(\\sqrt{1-x^2}\\right)}{K(x)}}, \\tag{1}\n$$</span> where <span class=\"math-container\">$q_k$</span> is the elliptic nome of modulus <span class=\"math-container\">$x$</span> and <span class=\"math-container\">$K(x)$</span> the complete elliptic integral of the first kind.\nOnce this is known one may compute\n<span class=\"math-container\">$$\n\\begin{aligned}\n\\det(I - M_\\infty^2) &amp;= \\prod_{p=0}^\\infty\\left(1-\\left(\\frac{2}{q_x^{p+\\frac12}+q_x^{-p-\\frac12}}\\right)^2\\right) \\\\\n&amp;=\\prod_{p=0}^\\infty\\left(\\frac{1-q_x^{2p+1}}{1+q_x^{2p+1}}\\right)^2 = \\frac{\\theta_4(0,q_x)}{\\theta_3(0,q_x)}=\\sqrt[4]{1-x^2},\n\\end{aligned}\n$$</span>\nwhere the last two equalities follow from standard properties of the <a href=\"https://mathworld.wolfram.com/JacobiThetaFunctions.html\" rel=\"nofollow noreferrer\">Jacobi theta functions <span class=\"math-container\">$\\theta_i$</span></a>.</p>\n<p>To see how the eigenvalue decomposition can be obtained, it is convenient to look at the Dirichlet space <span class=\"math-container\">$\\mathcal{D}$</span> of complex analytic functions <span class=\"math-container\">$f$</span> on the open unit disk that vanish at <span class=\"math-container\">$0$</span> and that have finite norm with respect to the Dirichlet inner product\n<span class=\"math-container\">$$\\langle f,g\\rangle_{\\mathcal{D}} = \\frac{1}{\\pi} \\int_{|z|&lt;1} \\overline{f'(z)}g'(z) \\mathrm{d}^2z = \\sum_{n=1}^\\infty n\\,\\overline{[z^n]f(z)}\\,[z^n]g(z).$$</span>\nA basis is given by the monomials <span class=\"math-container\">$(e_n:=z^n)_{n\\geq 1}$</span>, which satisfy <span class=\"math-container\">$\\langle e_n,e_m\\rangle_{\\mathcal{D}}=n\\, \\delta_{n,m}$</span>. Let us also introduce the bounded linear operator on <span class=\"math-container\">$\\mathcal{D}$</span> given by\n<span class=\"math-container\">$$\\mathbf{R}_x f = f \\circ \\psi_x, \\qquad \\psi_x(z) = \\frac{1-\\sqrt{1-x z^2}}{\\sqrt{x}\\,z}.$$</span>\nBy power series expansion one may check that\n<span class=\"math-container\">$$ \\mathbf{R}_x e_p = \\sum_{\\ell=p}^\\infty \\left(\\frac{x}{4}\\right)^{\\ell/2} \\frac{p}{\\ell}\\binom{\\ell}{(\\ell+p)/2}\\, 1_{\\{\\ell+p\\text{ even}\\}} e_\\ell$$</span>\nand the adjoint with respect to <span class=\"math-container\">$\\langle \\cdot,\\cdot\\rangle_{\\mathcal{D}}$</span> is determined by\n<span class=\"math-container\">$$ \\mathbf{R}_x^\\dagger e_p = \\sum_{\\ell=1}^p \\left(\\frac{x}{4}\\right)^{p/2} \\binom{p}{(\\ell+p)/2}\\, 1_{\\{\\ell+p\\text{ even}\\}} e_\\ell.$$</span>\nIn particular, <span class=\"math-container\">$\\mathbf{R}_x\\mathbf{R}_x^\\dagger$</span> is a self-adjoint operator that preserves the even and odd functions in <span class=\"math-container\">$\\mathcal{D}$</span>.\nUp to a factor of <span class=\"math-container\">$2$</span> its matrix elements on the odd monomials are precisely <span class=\"math-container\">$[M_\\infty]_{i,j}$</span>:\n<span class=\"math-container\">$$\n\\begin{aligned}\n2\\mathbf{R}_x\\mathbf{R}_x^\\dagger e_{2i+1} &amp;= \\sum_{j=0}^\\infty \\sum_{\\ell=0}^{\\min(i,j)} 2\\left(\\frac{x}{4}\\right)^{i+j+1} \\binom{2i+1}{i+\\ell+1}\\frac{2\\ell+1}{2j+1} \\binom{2j+1}{j+\\ell+1} e_{2j+1}\\\\\n&amp;= \\sum_{j=0}^\\infty\\frac{2i+1}{2(i+j+1)}\\binom{i-1/2}i\\binom{j-1/2}jx^{i+j+1}e_{2j+1}\\\\\n&amp;= \\sum_{j=0}^\\infty [M_\\infty]_{i,j} e_{2j+1}.\n\\end{aligned}\n$$</span>\nIn <em>T. Budd</em>, <a href=\"http://dx.doi.org/10.1016/j.jcta.2019.105191\" rel=\"nofollow noreferrer\"><strong>Winding of simple walks on the square lattice</strong></a>, <a href=\"https://arxiv.org/abs/1709.04042\" rel=\"nofollow noreferrer\">arXiv:1709.04042</a>, Section 2.1, I examined a closely related operator <span class=\"math-container\">$\\mathbf{J}_x = \\mathbf{R}_x^\\dagger\\mathbf{R}_x$</span>, whose matrix elements count diagonal walks on <span class=\"math-container\">$\\mathbb{Z}^2$</span> starting on the positive <span class=\"math-container\">$x$</span>-axis and hitting the <span class=\"math-container\">$y$</span>-axis at prescribed height. By a convenient elliptic parametrization of the disc one can determine all eigenvectors of <span class=\"math-container\">$\\mathbf{J}_x$</span>. Since <span class=\"math-container\">$\\mathbf{R}_x$</span> is injective, <span class=\"math-container\">$\\mathbf{R}_x\\mathbf{R}_x^\\dagger$</span> has the same eigenvalue decomposition (after applying <span class=\"math-container\">$\\mathbf{R}_x$</span> to the eigenvectors of <span class=\"math-container\">$\\mathbf{J}_x$</span>) given by Proposition 9, of which the eigenvectors with odd label <span class=\"math-container\">$m=1,3,5,\\ldots$</span>\nspan the odd functions in <span class=\"math-container\">$\\mathcal{D}$</span>. The corresponding eigenvalues are precisely the ones given above in (1).</p>\n<p><strong>Added (22 Jan):</strong> One may also check the formula for <span class=\"math-container\">$\\det(I\\pm M_\\infty)$</span> proposed in the <a href=\"https://mathoverflow.net/a/325953/47484\">post by Hucht</a>. Using the <a href=\"https://dlmf.nist.gov/20.4\" rel=\"nofollow noreferrer\">theta function product formulas</a> we find\n<span class=\"math-container\">$$\n\\det(I-M_\\infty) = \\prod_{p=0}^\\infty \\frac{(1-q_x^{p+\\frac12})^2}{1+q^{2p+1}} = \\left(\\frac{\\theta_2(0,\\sqrt{q_x})}{\\theta_1'(0,\\sqrt{q_x})}\\right)^{\\frac14} \\frac{\\theta_4(0,\\sqrt{q_x})}{\\theta_3(0,q_x)^{\\frac12}}.\n$$</span>\nWith the help of <span class=\"math-container\">$\\theta_1'(0,\\sqrt{q_x}) = \\theta_2(0,\\sqrt{q_x})\\theta_3(0,\\sqrt{q_x})\\theta_4(0,\\sqrt{q_x})$</span> and\n<span class=\"math-container\">$$\n\\theta_3(0,\\sqrt{q_x})^2 = (1+x)\\theta_3(0,q_x)^2,\\quad \\theta_4(0,\\sqrt{q_x})^2 = (1-x)\\theta_3(0,q_x)^2\n$$</span>\nthis yields\n<span class=\"math-container\">$$\n\\det(I- M_\\infty) = \\left(\\frac{(1-x)^3}{1+x}\\right)^{1/8}.\n$$</span>\nSimilarly\n<span class=\"math-container\">$$\n\\det(I-M_\\infty) = \\prod_{p=0}^\\infty \\frac{(1+q_x^{p+\\frac12})^2}{1+q^{2p+1}} = \\left(\\frac{\\theta_2(0,\\sqrt{q_x})}{\\theta_1'(0,\\sqrt{q_x})}\\right)^{\\frac14} \\frac{\\theta_3(0,\\sqrt{q_x})}{\\theta_3(0,q_x)^{\\frac12}}= \\left(\\frac{(1+x)^3}{1-x}\\right)^{1/8}.\n$$</span></p>\n" }, { "answer_id": 325953, "score": 2, "text": "<p>Based on your comment to my posting \"<a href=\"https://mathoverflow.net/questions/325886\">Conjecture for a certain Cauchy-type determinant</a>\", I found that in addition to the formula above (I write <span class=\"math-container\">$q$</span> instead of <span class=\"math-container\">$x$</span>, as all this is related to <span class=\"math-container\">$q$</span>-series, <span class=\"math-container\">$q$</span>-products, and Jacobi elliptic functions)\n<span class=\"math-container\">$$\n\\det[\\mathbf I_\\infty - \\mathbf M_\\infty(q) \\, \\mathbf M_\\infty(q)]\n=(1-q^2)^{1/4} \\qquad(1)\n$$</span>\nanother identity holds, which is closer to my case, namely\n<span class=\"math-container\">$$\n\\det[\\mathbf I_\\infty + \\mathbf M_\\infty^T(q) \\,\\mathbf M_\\infty(q)]\n=(1-q^2)^{-1/4}. \\qquad(2)\n$$</span>\nPlaying around a little further using my favorite tool <em>Mathematica</em>, I found the two \"roots\" of (1),\n<span class=\"math-container\">$$\n\\det[\\mathbf I_\\infty \\pm \\mathbf M_\\infty(q)] = \n\\left[\\frac{(1 \\pm q)^{3}}{1 \\mp q}\\right]^{1/8}.\n$$</span>\nI think that the resulting identity \n<span class=\"math-container\">$$\n\\det[\\mathbf I_\\infty + \\mathbf M_\\infty(q)] = \n\\det[\\mathbf I_\\infty - \\mathbf M_\\infty(-q)]\n$$</span>\ncan be proven quite easily by analyzing the series expansions.</p>\n\n<p>As <span class=\"math-container\">$\\mathbf M_\\infty$</span> and <span class=\"math-container\">$\\mathbf M_\\infty^T$</span> do not commute, (2) cannot easily be factored. \nHowever, maybe this helps a bit.</p>\n\n<p>Another observation is, that my problem <a href=\"https://mathoverflow.net/questions/325886\">325886</a> is an infinite product of this one, as\n<span class=\"math-container\">$$\n\\det[\\mathbf I_\\infty + \\mathbf X_\\infty^T(q) \\,\\mathbf X_\\infty(q)]\n= \\prod_{k=1}^{\\infty} \n\\det[\\mathbf I_\\infty + \\mathbf M_\\infty^T(q^k) \\,\\mathbf M_\\infty(q^k)].\n$$</span>\nSo they are definitely related. Obviously, all these expressions are matrix versions of <span class=\"math-container\">$q$</span>-product formulas. For further observations, see my updated posting <a href=\"https://mathoverflow.net/questions/325886\">325886</a>.</p>\n" } ]
Question: A neat evaluation of an infinite matrix? Answer 1: <p>The claimed expression for the determinant indeed holds. In fact, one may explicitly determine the eigenvalue decomposition of the infinite matrix <span class="math-container">$M_\infty$</span> viewed as a self-adjoint operator on a Hilbert space. Its eigenvalues are <span class="math-container">$$ \left(\frac{2}{q_x^{p+\frac12}+q_x^{-p-\frac12}}\right)_{p=0}^\infty,\qquad q_k = e^{-\pi\frac{K\left(\sqrt{1-x^2}\right)}{K(x)}}, \tag{1} $$</span> where <span class="math-container">$q_k$</span> is the elliptic nome of modulus <span class="math-container">$x$</span> and <span class="math-container">$K(x)$</span> the complete elliptic integral of the first kind. Once this is known one may compute <span class="math-container">$$ \begin{aligned} \det(I - M_\infty^2) &amp;= \prod_{p=0}^\infty\left(1-\left(\frac{2}{q_x^{p+\frac12}+q_x^{-p-\frac12}}\right)^2\right) \\ &amp;=\prod_{p=0}^\infty\left(\frac{1-q_x^{2p+1}}{1+q_x^{2p+1}}\right)^2 = \frac{\theta_4(0,q_x)}{\theta_3(0,q_x)}=\sqrt[4]{1-x^2}, \end{aligned} $$</span> where the last two equalities follow from standard properties of the <a href="https://mathworld.wolfram.com/JacobiThetaFunctions.html" rel="nofollow noreferrer">Jacobi theta functions <span class="math-container">$\theta_i$</span></a>.</p> <p>To see how the eigenvalue decomposition can be obtained, it is convenient to look at the Dirichlet space <span class="math-container">$\mathcal{D}$</span> of complex analytic functions <span class="math-container">$f$</span> on the open unit disk that vanish at <span class="math-container">$0$</span> and that have finite norm with respect to the Dirichlet inner product <span class="math-container">$$\langle f,g\rangle_{\mathcal{D}} = \frac{1}{\pi} \int_{|z|&lt;1} \overline{f'(z)}g'(z) \mathrm{d}^2z = \sum_{n=1}^\infty n\,\overline{[z^n]f(z)}\,[z^n]g(z).$$</span> A basis is given by the monomials <span class="math-container">$(e_n:=z^n)_{n\geq 1}$</span>, which satisfy <span class="math-container">$\langle e_n,e_m\rangle_{\mathcal{D}}=n\, \delta_{n,m}$</span>. Let us also introduce the bounded linear operator on <span class="math-container">$\mathcal{D}$</span> given by <span class="math-container">$$\mathbf{R}_x f = f \circ \psi_x, \qquad \psi_x(z) = \frac{1-\sqrt{1-x z^2}}{\sqrt{x}\,z}.$$</span> By power series expansion one may check that <span class="math-container">$$ \mathbf{R}_x e_p = \sum_{\ell=p}^\infty \left(\frac{x}{4}\right)^{\ell/2} \frac{p}{\ell}\binom{\ell}{(\ell+p)/2}\, 1_{\{\ell+p\text{ even}\}} e_\ell$$</span> and the adjoint with respect to <span class="math-container">$\langle \cdot,\cdot\rangle_{\mathcal{D}}$</span> is determined by <span class="math-container">$$ \mathbf{R}_x^\dagger e_p = \sum_{\ell=1}^p \left(\frac{x}{4}\right)^{p/2} \binom{p}{(\ell+p)/2}\, 1_{\{\ell+p\text{ even}\}} e_\ell.$$</span> In particular, <span class="math-container">$\mathbf{R}_x\mathbf{R}_x^\dagger$</span> is a self-adjoint operator that preserves the even and odd functions in <span class="math-container">$\mathcal{D}$</span>. Up to a factor of <span class="math-container">$2$</span> its matrix elements on the odd monomials are precisely <span class="math-container">$[M_\infty]_{i,j}$</span>: <span class="math-container">$$ \begin{aligned} 2\mathbf{R}_x\mathbf{R}_x^\dagger e_{2i+1} &amp;= \sum_{j=0}^\infty \sum_{\ell=0}^{\min(i,j)} 2\left(\frac{x}{4}\right)^{i+j+1} \binom{2i+1}{i+\ell+1}\frac{2\ell+1}{2j+1} \binom{2j+1}{j+\ell+1} e_{2j+1}\\ &amp;= \sum_{j=0}^\infty\frac{2i+1}{2(i+j+1)}\binom{i-1/2}i\binom{j-1/2}jx^{i+j+1}e_{2j+1}\\ &amp;= \sum_{j=0}^\infty [M_\infty]_{i,j} e_{2j+1}. \end{aligned} $$</span> In <em>T. Budd</em>, <a href="http://dx.doi.org/10.1016/j.jcta.2019.105191" rel="nofollow noreferrer"><strong>Winding of simple walks on the square lattice</strong></a>, <a href="https://arxiv.org/abs/1709.04042" rel="nofollow noreferrer">arXiv:1709.04042</a>, Section 2.1, I examined a closely related operator <span class="math-container">$\mathbf{J}_x = \mathbf{R}_x^\dagger\mathbf{R}_x$</span>, whose matrix elements count diagonal walks on <span class="math-container">$\mathbb{Z}^2$</span> starting on the positive <span class="math-container">$x$</span>-axis and hitting the <span class="math-container">$y$</span>-axis at prescribed height. By a convenient elliptic parametrization of the disc one can determine all eigenvectors of <span class="math-container">$\mathbf{J}_x$</span>. Since <span class="math-container">$\mathbf{R}_x$</span> is injective, <span class="math-container">$\mathbf{R}_x\mathbf{R}_x^\dagger$</span> has the same eigenvalue decomposition (after applying <span class="math-container">$\mathbf{R}_x$</span> to the eigenvectors of <span class="math-container">$\mathbf{J}_x$</span>) given by Proposition 9, of which the eigenvectors with odd label <span class="math-container">$m=1,3,5,\ldots$</span> span the odd functions in <span class="math-container">$\mathcal{D}$</span>. The corresponding eigenvalues are precisely the ones given above in (1).</p> <p><strong>Added (22 Jan):</strong> One may also check the formula for <span class="math-container">$\det(I\pm M_\infty)$</span> proposed in the <a href="https://mathoverflow.net/a/325953/47484">post by Hucht</a>. Using the <a href="https://dlmf.nist.gov/20.4" rel="nofollow noreferrer">theta function product formulas</a> we find <span class="math-container">$$ \det(I-M_\infty) = \prod_{p=0}^\infty \frac{(1-q_x^{p+\frac12})^2}{1+q^{2p+1}} = \left(\frac{\theta_2(0,\sqrt{q_x})}{\theta_1'(0,\sqrt{q_x})}\right)^{\frac14} \frac{\theta_4(0,\sqrt{q_x})}{\theta_3(0,q_x)^{\frac12}}. $$</span> With the help of <span class="math-container">$\theta_1'(0,\sqrt{q_x}) = \theta_2(0,\sqrt{q_x})\theta_3(0,\sqrt{q_x})\theta_4(0,\sqrt{q_x})$</span> and <span class="math-container">$$ \theta_3(0,\sqrt{q_x})^2 = (1+x)\theta_3(0,q_x)^2,\quad \theta_4(0,\sqrt{q_x})^2 = (1-x)\theta_3(0,q_x)^2 $$</span> this yields <span class="math-container">$$ \det(I- M_\infty) = \left(\frac{(1-x)^3}{1+x}\right)^{1/8}. $$</span> Similarly <span class="math-container">$$ \det(I-M_\infty) = \prod_{p=0}^\infty \frac{(1+q_x^{p+\frac12})^2}{1+q^{2p+1}} = \left(\frac{\theta_2(0,\sqrt{q_x})}{\theta_1'(0,\sqrt{q_x})}\right)^{\frac14} \frac{\theta_3(0,\sqrt{q_x})}{\theta_3(0,q_x)^{\frac12}}= \left(\frac{(1+x)^3}{1-x}\right)^{1/8}. $$</span></p> Answer 2: <p>Based on your comment to my posting "<a href="https://mathoverflow.net/questions/325886">Conjecture for a certain Cauchy-type determinant</a>", I found that in addition to the formula above (I write <span class="math-container">$q$</span> instead of <span class="math-container">$x$</span>, as all this is related to <span class="math-container">$q$</span>-series, <span class="math-container">$q$</span>-products, and Jacobi elliptic functions) <span class="math-container">$$ \det[\mathbf I_\infty - \mathbf M_\infty(q) \, \mathbf M_\infty(q)] =(1-q^2)^{1/4} \qquad(1) $$</span> another identity holds, which is closer to my case, namely <span class="math-container">$$ \det[\mathbf I_\infty + \mathbf M_\infty^T(q) \,\mathbf M_\infty(q)] =(1-q^2)^{-1/4}. \qquad(2) $$</span> Playing around a little further using my favorite tool <em>Mathematica</em>, I found the two "roots" of (1), <span class="math-container">$$ \det[\mathbf I_\infty \pm \mathbf M_\infty(q)] = \left[\frac{(1 \pm q)^{3}}{1 \mp q}\right]^{1/8}. $$</span> I think that the resulting identity <span class="math-container">$$ \det[\mathbf I_\infty + \mathbf M_\infty(q)] = \det[\mathbf I_\infty - \mathbf M_\infty(-q)] $$</span> can be proven quite easily by analyzing the series expansions.</p> <p>As <span class="math-container">$\mathbf M_\infty$</span> and <span class="math-container">$\mathbf M_\infty^T$</span> do not commute, (2) cannot easily be factored. However, maybe this helps a bit.</p> <p>Another observation is, that my problem <a href="https://mathoverflow.net/questions/325886">325886</a> is an infinite product of this one, as <span class="math-container">$$ \det[\mathbf I_\infty + \mathbf X_\infty^T(q) \,\mathbf X_\infty(q)] = \prod_{k=1}^{\infty} \det[\mathbf I_\infty + \mathbf M_\infty^T(q^k) \,\mathbf M_\infty(q^k)]. $$</span> So they are definitely related. Obviously, all these expressions are matrix versions of <span class="math-container">$q$</span>-product formulas. For further observations, see my updated posting <a href="https://mathoverflow.net/questions/325886">325886</a>.</p> Answer 3: <p>The claimed expression for the determinant indeed holds. In fact, one may explicitly determine the eigenvalue decomposition of the infinite matrix <span class="math-container">$M_\infty$</span> viewed as a self-adjoint operator on a Hilbert space. Its eigenvalues are <span class="math-container">$$ \left(\frac{2}{q_x^{p+\frac12}+q_x^{-p-\frac12}}\right)_{p=0}^\infty,\qquad q_k = e^{-\pi\frac{K\left(\sqrt{1-x^2}\right)}{K(x)}}, \tag{1} $$</span> where <span class="math-container">$q_k$</span> is the elliptic nome of modulus <span class="math-container">$x$</span> and <span class="math-container">$K(x)$</span> the complete elliptic integral of the first kind. Once this is known one may compute <span class="math-container">$$ \begin{aligned} \det(I - M_\infty^2) &amp;= \prod_{p=0}^\infty\left(1-\left(\frac{2}{q_x^{p+\frac12}+q_x^{-p-\frac12}}\right)^2\right) \\ &amp;=\prod_{p=0}^\infty\left(\frac{1-q_x^{2p+1}}{1+q_x^{2p+1}}\right)^2 = \frac{\theta_4(0,q_x)}{\theta_3(0,q_x)}=\sqrt[4]{1-x^2}, \end{aligned} $$</span> where the last two equalities follow from standard properties of the <a href="https://mathworld.wolfram.com/JacobiThetaFunctions.html" rel="nofollow noreferrer">Jacobi theta functions <span class="math-container">$\theta_i$</span></a>.</p> <p>To see how the eigenvalue decomposition can be obtained, it is convenient to look at the Dirichlet space <span class="math-container">$\mathcal{D}$</span> of complex analytic functions <span class="math-container">$f$</span> on the open unit disk that vanish at <span class="math-container">$0$</span> and that have finite norm with respect to the Dirichlet inner product <span class="math-container">$$\langle f,g\rangle_{\mathcal{D}} = \frac{1}{\pi} \int_{|z|&lt;1} \overline{f'(z)}g'(z) \mathrm{d}^2z = \sum_{n=1}^\infty n\,\overline{[z^n]f(z)}\,[z^n]g(z).$$</span> A basis is given by the monomials <span class="math-container">$(e_n:=z^n)_{n\geq 1}$</span>, which satisfy <span class="math-container">$\langle e_n,e_m\rangle_{\mathcal{D}}=n\, \delta_{n,m}$</span>. Let us also introduce the bounded linear operator on <span class="math-container">$\mathcal{D}$</span> given by <span class="math-container">$$\mathbf{R}_x f = f \circ \psi_x, \qquad \psi_x(z) = \frac{1-\sqrt{1-x z^2}}{\sqrt{x}\,z}.$$</span> By power series expansion one may check that <span class="math-container">$$ \mathbf{R}_x e_p = \sum_{\ell=p}^\infty \left(\frac{x}{4}\right)^{\ell/2} \frac{p}{\ell}\binom{\ell}{(\ell+p)/2}\, 1_{\{\ell+p\text{ even}\}} e_\ell$$</span> and the adjoint with respect to <span class="math-container">$\langle \cdot,\cdot\rangle_{\mathcal{D}}$</span> is determined by <span class="math-container">$$ \mathbf{R}_x^\dagger e_p = \sum_{\ell=1}^p \left(\frac{x}{4}\right)^{p/2} \binom{p}{(\ell+p)/2}\, 1_{\{\ell+p\text{ even}\}} e_\ell.$$</span> In particular, <span class="math-container">$\mathbf{R}_x\mathbf{R}_x^\dagger$</span> is a self-adjoint operator that preserves the even and odd functions in <span class="math-container">$\mathcal{D}$</span>. Up to a factor of <span class="math-container">$2$</span> its matrix elements on the odd monomials are precisely <span class="math-container">$[M_\infty]_{i,j}$</span>: <span class="math-container">$$ \begin{aligned} 2\mathbf{R}_x\mathbf{R}_x^\dagger e_{2i+1} &amp;= \sum_{j=0}^\infty \sum_{\ell=0}^{\min(i,j)} 2\left(\frac{x}{4}\right)^{i+j+1} \binom{2i+1}{i+\ell+1}\frac{2\ell+1}{2j+1} \binom{2j+1}{j+\ell+1} e_{2j+1}\\ &amp;= \sum_{j=0}^\infty\frac{2i+1}{2(i+j+1)}\binom{i-1/2}i\binom{j-1/2}jx^{i+j+1}e_{2j+1}\\ &amp;= \sum_{j=0}^\infty [M_\infty]_{i,j} e_{2j+1}. \end{aligned} $$</span> In <em>T. Budd</em>, <a href="http://dx.doi.org/10.1016/j.jcta.2019.105191" rel="nofollow noreferrer"><strong>Winding of simple walks on the square lattice</strong></a>, <a href="https://arxiv.org/abs/1709.04042" rel="nofollow noreferrer">arXiv:1709.04042</a>, Section 2.1, I examined a closely related operator <span class="math-container">$\mathbf{J}_x = \mathbf{R}_x^\dagger\mathbf{R}_x$</span>, whose matrix elements count diagonal walks on <span class="math-container">$\mathbb{Z}^2$</span> starting on the positive <span class="math-container">$x$</span>-axis and hitting the <span class="math-container">$y$</span>-axis at prescribed height. By a convenient elliptic parametrization of the disc one can determine all eigenvectors of <span class="math-container">$\mathbf{J}_x$</span>. Since <span class="math-container">$\mathbf{R}_x$</span> is injective, <span class="math-container">$\mathbf{R}_x\mathbf{R}_x^\dagger$</span> has the same eigenvalue decomposition (after applying <span class="math-container">$\mathbf{R}_x$</span> to the eigenvectors of <span class="math-container">$\mathbf{J}_x$</span>) given by Proposition 9, of which the eigenvectors with odd label <span class="math-container">$m=1,3,5,\ldots$</span> span the odd functions in <span class="math-container">$\mathcal{D}$</span>. The corresponding eigenvalues are precisely the ones given above in (1).</p> <p><strong>Added (22 Jan):</strong> One may also check the formula for <span class="math-container">$\det(I\pm M_\infty)$</span> proposed in the <a href="https://mathoverflow.net/a/325953/47484">post by Hucht</a>. Using the <a href="https://dlmf.nist.gov/20.4" rel="nofollow noreferrer">theta function product formulas</a> we find <span class="math-container">$$ \det(I-M_\infty) = \prod_{p=0}^\infty \frac{(1-q_x^{p+\frac12})^2}{1+q^{2p+1}} = \left(\frac{\theta_2(0,\sqrt{q_x})}{\theta_1'(0,\sqrt{q_x})}\right)^{\frac14} \frac{\theta_4(0,\sqrt{q_x})}{\theta_3(0,q_x)^{\frac12}}. $$</span> With the help of <span class="math-container">$\theta_1'(0,\sqrt{q_x}) = \theta_2(0,\sqrt{q_x})\theta_3(0,\sqrt{q_x})\theta_4(0,\sqrt{q_x})$</span> and <span class="math-container">$$ \theta_3(0,\sqrt{q_x})^2 = (1+x)\theta_3(0,q_x)^2,\quad \theta_4(0,\sqrt{q_x})^2 = (1-x)\theta_3(0,q_x)^2 $$</span> this yields <span class="math-container">$$ \det(I- M_\infty) = \left(\frac{(1-x)^3}{1+x}\right)^{1/8}. $$</span> Similarly <span class="math-container">$$ \det(I-M_\infty) = \prod_{p=0}^\infty \frac{(1+q_x^{p+\frac12})^2}{1+q^{2p+1}} = \left(\frac{\theta_2(0,\sqrt{q_x})}{\theta_1'(0,\sqrt{q_x})}\right)^{\frac14} \frac{\theta_3(0,\sqrt{q_x})}{\theta_3(0,q_x)^{\frac12}}= \left(\frac{(1+x)^3}{1-x}\right)^{1/8}. $$</span></p> Answer 4: <p>Based on your comment to my posting "<a href="https://mathoverflow.net/questions/325886">Conjecture for a certain Cauchy-type determinant</a>", I found that in addition to the formula above (I write <span class="math-container">$q$</span> instead of <span class="math-container">$x$</span>, as all this is related to <span class="math-container">$q$</span>-series, <span class="math-container">$q$</span>-products, and Jacobi elliptic functions) <span class="math-container">$$ \det[\mathbf I_\infty - \mathbf M_\infty(q) \, \mathbf M_\infty(q)] =(1-q^2)^{1/4} \qquad(1) $$</span> another identity holds, which is closer to my case, namely <span class="math-container">$$ \det[\mathbf I_\infty + \mathbf M_\infty^T(q) \,\mathbf M_\infty(q)] =(1-q^2)^{-1/4}. \qquad(2) $$</span> Playing around a little further using my favorite tool <em>Mathematica</em>, I found the two "roots" of (1), <span class="math-container">$$ \det[\mathbf I_\infty \pm \mathbf M_\infty(q)] = \left[\frac{(1 \pm q)^{3}}{1 \mp q}\right]^{1/8}. $$</span> I think that the resulting identity <span class="math-container">$$ \det[\mathbf I_\infty + \mathbf M_\infty(q)] = \det[\mathbf I_\infty - \mathbf M_\infty(-q)] $$</span> can be proven quite easily by analyzing the series expansions.</p> <p>As <span class="math-container">$\mathbf M_\infty$</span> and <span class="math-container">$\mathbf M_\infty^T$</span> do not commute, (2) cannot easily be factored. However, maybe this helps a bit.</p> <p>Another observation is, that my problem <a href="https://mathoverflow.net/questions/325886">325886</a> is an infinite product of this one, as <span class="math-container">$$ \det[\mathbf I_\infty + \mathbf X_\infty^T(q) \,\mathbf X_\infty(q)] = \prod_{k=1}^{\infty} \det[\mathbf I_\infty + \mathbf M_\infty^T(q^k) \,\mathbf M_\infty(q^k)]. $$</span> So they are definitely related. Obviously, all these expressions are matrix versions of <span class="math-container">$q$</span>-product formulas. For further observations, see my updated posting <a href="https://mathoverflow.net/questions/325886">325886</a>.</p> Answer 5: <p>The claimed expression for the determinant indeed holds. In fact, one may explicitly determine the eigenvalue decomposition of the infinite matrix <span class="math-container">$M_\infty$</span> viewed as a self-adjoint operator on a Hilbert space. Its eigenvalues are <span class="math-container">$$ \left(\frac{2}{q_x^{p+\frac12}+q_x^{-p-\frac12}}\right)_{p=0}^\infty,\qquad q_k = e^{-\pi\frac{K\left(\sqrt{1-x^2}\right)}{K(x)}}, \tag{1} $$</span> where <span class="math-container">$q_k$</span> is the elliptic nome of modulus <span class="math-container">$x$</span> and <span class="math-container">$K(x)$</span> the complete elliptic integral of the first kind. Once this is known one may compute <span class="math-container">$$ \begin{aligned} \det(I - M_\infty^2) &amp;= \prod_{p=0}^\infty\left(1-\left(\frac{2}{q_x^{p+\frac12}+q_x^{-p-\frac12}}\right)^2\right) \\ &amp;=\prod_{p=0}^\infty\left(\frac{1-q_x^{2p+1}}{1+q_x^{2p+1}}\right)^2 = \frac{\theta_4(0,q_x)}{\theta_3(0,q_x)}=\sqrt[4]{1-x^2}, \end{aligned} $$</span> where the last two equalities follow from standard properties of the <a href="https://mathworld.wolfram.com/JacobiThetaFunctions.html" rel="nofollow noreferrer">Jacobi theta functions <span class="math-container">$\theta_i$</span></a>.</p> <p>To see how the eigenvalue decomposition can be obtained, it is convenient to look at the Dirichlet space <span class="math-container">$\mathcal{D}$</span> of complex analytic functions <span class="math-container">$f$</span> on the open unit disk that vanish at <span class="math-container">$0$</span> and that have finite norm with respect to the Dirichlet inner product <span class="math-container">$$\langle f,g\rangle_{\mathcal{D}} = \frac{1}{\pi} \int_{|z|&lt;1} \overline{f'(z)}g'(z) \mathrm{d}^2z = \sum_{n=1}^\infty n\,\overline{[z^n]f(z)}\,[z^n]g(z).$$</span> A basis is given by the monomials <span class="math-container">$(e_n:=z^n)_{n\geq 1}$</span>, which satisfy <span class="math-container">$\langle e_n,e_m\rangle_{\mathcal{D}}=n\, \delta_{n,m}$</span>. Let us also introduce the bounded linear operator on <span class="math-container">$\mathcal{D}$</span> given by <span class="math-container">$$\mathbf{R}_x f = f \circ \psi_x, \qquad \psi_x(z) = \frac{1-\sqrt{1-x z^2}}{\sqrt{x}\,z}.$$</span> By power series expansion one may check that <span class="math-container">$$ \mathbf{R}_x e_p = \sum_{\ell=p}^\infty \left(\frac{x}{4}\right)^{\ell/2} \frac{p}{\ell}\binom{\ell}{(\ell+p)/2}\, 1_{\{\ell+p\text{ even}\}} e_\ell$$</span> and the adjoint with respect to <span class="math-container">$\langle \cdot,\cdot\rangle_{\mathcal{D}}$</span> is determined by <span class="math-container">$$ \mathbf{R}_x^\dagger e_p = \sum_{\ell=1}^p \left(\frac{x}{4}\right)^{p/2} \binom{p}{(\ell+p)/2}\, 1_{\{\ell+p\text{ even}\}} e_\ell.$$</span> In particular, <span class="math-container">$\mathbf{R}_x\mathbf{R}_x^\dagger$</span> is a self-adjoint operator that preserves the even and odd functions in <span class="math-container">$\mathcal{D}$</span>. Up to a factor of <span class="math-container">$2$</span> its matrix elements on the odd monomials are precisely <span class="math-container">$[M_\infty]_{i,j}$</span>: <span class="math-container">$$ \begin{aligned} 2\mathbf{R}_x\mathbf{R}_x^\dagger e_{2i+1} &amp;= \sum_{j=0}^\infty \sum_{\ell=0}^{\min(i,j)} 2\left(\frac{x}{4}\right)^{i+j+1} \binom{2i+1}{i+\ell+1}\frac{2\ell+1}{2j+1} \binom{2j+1}{j+\ell+1} e_{2j+1}\\ &amp;= \sum_{j=0}^\infty\frac{2i+1}{2(i+j+1)}\binom{i-1/2}i\binom{j-1/2}jx^{i+j+1}e_{2j+1}\\ &amp;= \sum_{j=0}^\infty [M_\infty]_{i,j} e_{2j+1}. \end{aligned} $$</span> In <em>T. Budd</em>, <a href="http://dx.doi.org/10.1016/j.jcta.2019.105191" rel="nofollow noreferrer"><strong>Winding of simple walks on the square lattice</strong></a>, <a href="https://arxiv.org/abs/1709.04042" rel="nofollow noreferrer">arXiv:1709.04042</a>, Section 2.1, I examined a closely related operator <span class="math-container">$\mathbf{J}_x = \mathbf{R}_x^\dagger\mathbf{R}_x$</span>, whose matrix elements count diagonal walks on <span class="math-container">$\mathbb{Z}^2$</span> starting on the positive <span class="math-container">$x$</span>-axis and hitting the <span class="math-container">$y$</span>-axis at prescribed height. By a convenient elliptic parametrization of the disc one can determine all eigenvectors of <span class="math-container">$\mathbf{J}_x$</span>. Since <span class="math-container">$\mathbf{R}_x$</span> is injective, <span class="math-container">$\mathbf{R}_x\mathbf{R}_x^\dagger$</span> has the same eigenvalue decomposition (after applying <span class="math-container">$\mathbf{R}_x$</span> to the eigenvectors of <span class="math-container">$\mathbf{J}_x$</span>) given by Proposition 9, of which the eigenvectors with odd label <span class="math-container">$m=1,3,5,\ldots$</span> span the odd functions in <span class="math-container">$\mathcal{D}$</span>. The corresponding eigenvalues are precisely the ones given above in (1).</p> <p><strong>Added (22 Jan):</strong> One may also check the formula for <span class="math-container">$\det(I\pm M_\infty)$</span> proposed in the <a href="https://mathoverflow.net/a/325953/47484">post by Hucht</a>. Using the <a href="https://dlmf.nist.gov/20.4" rel="nofollow noreferrer">theta function product formulas</a> we find <span class="math-container">$$ \det(I-M_\infty) = \prod_{p=0}^\infty \frac{(1-q_x^{p+\frac12})^2}{1+q^{2p+1}} = \left(\frac{\theta_2(0,\sqrt{q_x})}{\theta_1'(0,\sqrt{q_x})}\right)^{\frac14} \frac{\theta_4(0,\sqrt{q_x})}{\theta_3(0,q_x)^{\frac12}}. $$</span> With the help of <span class="math-container">$\theta_1'(0,\sqrt{q_x}) = \theta_2(0,\sqrt{q_x})\theta_3(0,\sqrt{q_x})\theta_4(0,\sqrt{q_x})$</span> and <span class="math-container">$$ \theta_3(0,\sqrt{q_x})^2 = (1+x)\theta_3(0,q_x)^2,\quad \theta_4(0,\sqrt{q_x})^2 = (1-x)\theta_3(0,q_x)^2 $$</span> this yields <span class="math-container">$$ \det(I- M_\infty) = \left(\frac{(1-x)^3}{1+x}\right)^{1/8}. $$</span> Similarly <span class="math-container">$$ \det(I-M_\infty) = \prod_{p=0}^\infty \frac{(1+q_x^{p+\frac12})^2}{1+q^{2p+1}} = \left(\frac{\theta_2(0,\sqrt{q_x})}{\theta_1'(0,\sqrt{q_x})}\right)^{\frac14} \frac{\theta_3(0,\sqrt{q_x})}{\theta_3(0,q_x)^{\frac12}}= \left(\frac{(1+x)^3}{1-x}\right)^{1/8}. $$</span></p> Answer 6: <p>Based on your comment to my posting "<a href="https://mathoverflow.net/questions/325886">Conjecture for a certain Cauchy-type determinant</a>", I found that in addition to the formula above (I write <span class="math-container">$q$</span> instead of <span class="math-container">$x$</span>, as all this is related to <span class="math-container">$q$</span>-series, <span class="math-container">$q$</span>-products, and Jacobi elliptic functions) <span class="math-container">$$ \det[\mathbf I_\infty - \mathbf M_\infty(q) \, \mathbf M_\infty(q)] =(1-q^2)^{1/4} \qquad(1) $$</span> another identity holds, which is closer to my case, namely <span class="math-container">$$ \det[\mathbf I_\infty + \mathbf M_\infty^T(q) \,\mathbf M_\infty(q)] =(1-q^2)^{-1/4}. \qquad(2) $$</span> Playing around a little further using my favorite tool <em>Mathematica</em>, I found the two "roots" of (1), <span class="math-container">$$ \det[\mathbf I_\infty \pm \mathbf M_\infty(q)] = \left[\frac{(1 \pm q)^{3}}{1 \mp q}\right]^{1/8}. $$</span> I think that the resulting identity <span class="math-container">$$ \det[\mathbf I_\infty + \mathbf M_\infty(q)] = \det[\mathbf I_\infty - \mathbf M_\infty(-q)] $$</span> can be proven quite easily by analyzing the series expansions.</p> <p>As <span class="math-container">$\mathbf M_\infty$</span> and <span class="math-container">$\mathbf M_\infty^T$</span> do not commute, (2) cannot easily be factored. However, maybe this helps a bit.</p> <p>Another observation is, that my problem <a href="https://mathoverflow.net/questions/325886">325886</a> is an infinite product of this one, as <span class="math-container">$$ \det[\mathbf I_\infty + \mathbf X_\infty^T(q) \,\mathbf X_\infty(q)] = \prod_{k=1}^{\infty} \det[\mathbf I_\infty + \mathbf M_\infty^T(q^k) \,\mathbf M_\infty(q^k)]. $$</span> So they are definitely related. Obviously, all these expressions are matrix versions of <span class="math-container">$q$</span>-product formulas. For further observations, see my updated posting <a href="https://mathoverflow.net/questions/325886">325886</a>.</p>
Cartier-Kostant-Milnor-Moore theorem
259,307
11
https://mathoverflow.net/questions/259307/cartier-kostant-milnor-moore-theorem
<p>If $k$ is an algebraically closed field of characteristic zero and $H$ is a cocommutative Hopf algebra, then $$ H \cong U(P(H)) \ltimes kG(H). $$ What happens if the field is not algebraically closed? Is the theorem still true or is there any counterexample? What about characteristic different from zero?</p>
[ { "answer_id": 259886, "score": 10, "text": "<p>When $k$ fails to be algebraically closed the theorem is false but the discrepancy can be understood in terms of Galois descent and so in principle understood in terms of Galois cohomology. </p>\n\n<p>Suppose $H$ is a cocommutative Hopf algebra over $k$, not algebraically closed but characteristic $0$. Then the classification theorem applies to the base change $H \\otimes_k \\bar{k}$ of $H$ to the algebraic closure, which must therefore take the form of the semidirect product of a universal enveloping algebra and a group algebra. So $H$ is a $k$-form of such a thing, but need not be such a thing itself.</p>\n\n<p>For example, suppose $H$ is finite-dimensional, so $H \\otimes_k \\bar{k}$ is a group algebra $\\bar{k}[\\Lambda]$ of a finite group $\\Lambda$. The most obvious $k$-form of $\\bar{k}[\\Lambda]$ is, of course, $k[\\Lambda]$, but other $k$-forms are possible. In this setting $H$ necessarily splits over a finite Galois extension $L$ in the sense that we must already have</p>\n\n<p>$$H \\otimes_k L \\cong L[\\Lambda].$$</p>\n\n<p>Now, the Galois group $G = \\text{Gal}(L/k)$ acts on the LHS by Hopf algebra automorphisms (over $k$), so also acts on the RHS by Hopf algebra automorphisms (over $k$), and hence acts on the group of grouplike elements $\\Lambda$ on the RHS. If we already had $H \\cong k[\\Lambda]$ then this action would be trivial, but in fact every possible action of $G$ on $\\Lambda$ occurs, and conjugacy classes of such actions classify the possible choices of $H$. Given such an action $\\rho : H \\to \\text{Aut}(\\Lambda)$ we can recover $H$ as</p>\n\n<p>$$H \\cong L[\\Lambda]^G$$</p>\n\n<p>where the action of $G$ on $L[\\Lambda]$ extends the action of $G$ on $\\Lambda$.</p>\n\n<p>Very explicitly, let $k = \\mathbb{R}, L = \\mathbb{C}$, and $\\Lambda = C_3$, where the Galois group $G = C_2$ acts via inversion. Then we have</p>\n\n<p>$$H \\otimes_{\\mathbb{R}} \\mathbb{C} \\cong \\mathbb{C}[C_3] \\cong \\mathbb{C}[x]/(x^3 - 1)$$</p>\n\n<p>where $x$ is grouplike but the action of complex conjugation sends $x$ to $x^{-1} = x^2$. The fixed point subalgebra $H$ has a basis given by the identity $1$ and the \"real and imaginary parts\"</p>\n\n<p>$$c = \\frac{x + x^{-1}}{2}$$\n$$s = \\frac{x - x^{-1}}{2i}$$</p>\n\n<p>(it helps to think of $x$ as secretly being $\\omega = e^{\\frac{2\\pi i}{3}}$ here) of $x$, with comultiplication given by the sine and cosine angle addition formulas</p>\n\n<p>$$\\Delta(c) = c \\otimes c - s \\otimes s$$\n$$\\Delta(s) = c \\otimes s + s \\otimes c.$$</p>\n\n<p>So neither $c$ nor $s$ is grouplike, and in fact $H$ has no nontrivial grouplike elements except the identity. In $H \\otimes_{\\mathbb{R}} \\mathbb{C}$ the nontrivial grouplike elements are $c + is$ and $c - is$. </p>\n" }, { "answer_id": 259613, "score": 6, "text": "<p>The case of irreducible, cocommutative Hopf algebras, over a field with $char(k)&gt; 0$, is discussed in Sweedler's textbook on Hopf algebras, Ch.$XIII$, sect. $13.2$. (See prop. $13.2.2$, $13.2.3$). </p>\n\n<p>For the case of cocommutative Hopf algebras, over an algebraically closed field $k$, with $char(k)\\geq 3$ you can have a look at: </p>\n\n<p><a href=\"https://www.semanticscholar.org/paper/Modules-of-Solvable-Innnitesimal-Groups-and-the-Farnsteiner-Voigt/a7ffd5538cd3e73320773c05ba4311e12e12905d\" rel=\"nofollow noreferrer\">Modules of solvable infinitesimal groups and the structure of representation-finite cocommutative Hopf algebras</a>, R.Farnsteiner, D.Voigt, Math. Proc. Cambridge Philos. Soc., v.127, p.441-459, 1999 </p>\n\n<p>and the references therein (among them, there is also an interesting older paper by the same authors discussing the case of cocommutative hopf algebras of finite representation type, over an algebraically closed field of $char(k)&gt;0$). </p>\n\n<p><strong>Edit (July 2018):</strong> Since user's Qiaochu Yuan answer, provides a form of counter-example to the theorem, for non-algebraically closed fields, i think it would be of some added value to indicate cases in which the Cartier-Kostant-Milnor-Moore theorem remains valid for non-algebraically closed fields: this happens if we take as an assumption that we are dealing with <a href=\"http://library.msri.org/books/Book43/files/andrus.pdf\" rel=\"nofollow noreferrer\">pointed</a> hopf algebras. So, the (somewhat more general) statement of the theorem goes like: </p>\n\n<blockquote>\n <p>If $H$ is a pointed, cocommutative hopf algebra over a field $k$ of characteristic $0$, we have that \n $$\nH\\cong U\\big(P(H)\\big)\\ltimes kG(H)\n$$\n where $P(H)$ is the lie algebra of the primitives, $U(.)$ its UEA and $G(H)$ the group of the group-likes. </p>\n</blockquote>\n\n<p>The algebraically closed field case (mentioned in the OP), may be considered a special case of the above, taking into account that cocommutative hopf algebras over algebraically closed fields, can be easily seen to be pointed. </p>\n" }, { "answer_id": 259886, "score": 10, "text": "<p>When $k$ fails to be algebraically closed the theorem is false but the discrepancy can be understood in terms of Galois descent and so in principle understood in terms of Galois cohomology. </p>\n\n<p>Suppose $H$ is a cocommutative Hopf algebra over $k$, not algebraically closed but characteristic $0$. Then the classification theorem applies to the base change $H \\otimes_k \\bar{k}$ of $H$ to the algebraic closure, which must therefore take the form of the semidirect product of a universal enveloping algebra and a group algebra. So $H$ is a $k$-form of such a thing, but need not be such a thing itself.</p>\n\n<p>For example, suppose $H$ is finite-dimensional, so $H \\otimes_k \\bar{k}$ is a group algebra $\\bar{k}[\\Lambda]$ of a finite group $\\Lambda$. The most obvious $k$-form of $\\bar{k}[\\Lambda]$ is, of course, $k[\\Lambda]$, but other $k$-forms are possible. In this setting $H$ necessarily splits over a finite Galois extension $L$ in the sense that we must already have</p>\n\n<p>$$H \\otimes_k L \\cong L[\\Lambda].$$</p>\n\n<p>Now, the Galois group $G = \\text{Gal}(L/k)$ acts on the LHS by Hopf algebra automorphisms (over $k$), so also acts on the RHS by Hopf algebra automorphisms (over $k$), and hence acts on the group of grouplike elements $\\Lambda$ on the RHS. If we already had $H \\cong k[\\Lambda]$ then this action would be trivial, but in fact every possible action of $G$ on $\\Lambda$ occurs, and conjugacy classes of such actions classify the possible choices of $H$. Given such an action $\\rho : H \\to \\text{Aut}(\\Lambda)$ we can recover $H$ as</p>\n\n<p>$$H \\cong L[\\Lambda]^G$$</p>\n\n<p>where the action of $G$ on $L[\\Lambda]$ extends the action of $G$ on $\\Lambda$.</p>\n\n<p>Very explicitly, let $k = \\mathbb{R}, L = \\mathbb{C}$, and $\\Lambda = C_3$, where the Galois group $G = C_2$ acts via inversion. Then we have</p>\n\n<p>$$H \\otimes_{\\mathbb{R}} \\mathbb{C} \\cong \\mathbb{C}[C_3] \\cong \\mathbb{C}[x]/(x^3 - 1)$$</p>\n\n<p>where $x$ is grouplike but the action of complex conjugation sends $x$ to $x^{-1} = x^2$. The fixed point subalgebra $H$ has a basis given by the identity $1$ and the \"real and imaginary parts\"</p>\n\n<p>$$c = \\frac{x + x^{-1}}{2}$$\n$$s = \\frac{x - x^{-1}}{2i}$$</p>\n\n<p>(it helps to think of $x$ as secretly being $\\omega = e^{\\frac{2\\pi i}{3}}$ here) of $x$, with comultiplication given by the sine and cosine angle addition formulas</p>\n\n<p>$$\\Delta(c) = c \\otimes c - s \\otimes s$$\n$$\\Delta(s) = c \\otimes s + s \\otimes c.$$</p>\n\n<p>So neither $c$ nor $s$ is grouplike, and in fact $H$ has no nontrivial grouplike elements except the identity. In $H \\otimes_{\\mathbb{R}} \\mathbb{C}$ the nontrivial grouplike elements are $c + is$ and $c - is$. </p>\n" }, { "answer_id": 259613, "score": 6, "text": "<p>The case of irreducible, cocommutative Hopf algebras, over a field with $char(k)&gt; 0$, is discussed in Sweedler's textbook on Hopf algebras, Ch.$XIII$, sect. $13.2$. (See prop. $13.2.2$, $13.2.3$). </p>\n\n<p>For the case of cocommutative Hopf algebras, over an algebraically closed field $k$, with $char(k)\\geq 3$ you can have a look at: </p>\n\n<p><a href=\"https://www.semanticscholar.org/paper/Modules-of-Solvable-Innnitesimal-Groups-and-the-Farnsteiner-Voigt/a7ffd5538cd3e73320773c05ba4311e12e12905d\" rel=\"nofollow noreferrer\">Modules of solvable infinitesimal groups and the structure of representation-finite cocommutative Hopf algebras</a>, R.Farnsteiner, D.Voigt, Math. Proc. Cambridge Philos. Soc., v.127, p.441-459, 1999 </p>\n\n<p>and the references therein (among them, there is also an interesting older paper by the same authors discussing the case of cocommutative hopf algebras of finite representation type, over an algebraically closed field of $char(k)&gt;0$). </p>\n\n<p><strong>Edit (July 2018):</strong> Since user's Qiaochu Yuan answer, provides a form of counter-example to the theorem, for non-algebraically closed fields, i think it would be of some added value to indicate cases in which the Cartier-Kostant-Milnor-Moore theorem remains valid for non-algebraically closed fields: this happens if we take as an assumption that we are dealing with <a href=\"http://library.msri.org/books/Book43/files/andrus.pdf\" rel=\"nofollow noreferrer\">pointed</a> hopf algebras. So, the (somewhat more general) statement of the theorem goes like: </p>\n\n<blockquote>\n <p>If $H$ is a pointed, cocommutative hopf algebra over a field $k$ of characteristic $0$, we have that \n $$\nH\\cong U\\big(P(H)\\big)\\ltimes kG(H)\n$$\n where $P(H)$ is the lie algebra of the primitives, $U(.)$ its UEA and $G(H)$ the group of the group-likes. </p>\n</blockquote>\n\n<p>The algebraically closed field case (mentioned in the OP), may be considered a special case of the above, taking into account that cocommutative hopf algebras over algebraically closed fields, can be easily seen to be pointed. </p>\n" }, { "answer_id": 259886, "score": 10, "text": "<p>When $k$ fails to be algebraically closed the theorem is false but the discrepancy can be understood in terms of Galois descent and so in principle understood in terms of Galois cohomology. </p>\n\n<p>Suppose $H$ is a cocommutative Hopf algebra over $k$, not algebraically closed but characteristic $0$. Then the classification theorem applies to the base change $H \\otimes_k \\bar{k}$ of $H$ to the algebraic closure, which must therefore take the form of the semidirect product of a universal enveloping algebra and a group algebra. So $H$ is a $k$-form of such a thing, but need not be such a thing itself.</p>\n\n<p>For example, suppose $H$ is finite-dimensional, so $H \\otimes_k \\bar{k}$ is a group algebra $\\bar{k}[\\Lambda]$ of a finite group $\\Lambda$. The most obvious $k$-form of $\\bar{k}[\\Lambda]$ is, of course, $k[\\Lambda]$, but other $k$-forms are possible. In this setting $H$ necessarily splits over a finite Galois extension $L$ in the sense that we must already have</p>\n\n<p>$$H \\otimes_k L \\cong L[\\Lambda].$$</p>\n\n<p>Now, the Galois group $G = \\text{Gal}(L/k)$ acts on the LHS by Hopf algebra automorphisms (over $k$), so also acts on the RHS by Hopf algebra automorphisms (over $k$), and hence acts on the group of grouplike elements $\\Lambda$ on the RHS. If we already had $H \\cong k[\\Lambda]$ then this action would be trivial, but in fact every possible action of $G$ on $\\Lambda$ occurs, and conjugacy classes of such actions classify the possible choices of $H$. Given such an action $\\rho : H \\to \\text{Aut}(\\Lambda)$ we can recover $H$ as</p>\n\n<p>$$H \\cong L[\\Lambda]^G$$</p>\n\n<p>where the action of $G$ on $L[\\Lambda]$ extends the action of $G$ on $\\Lambda$.</p>\n\n<p>Very explicitly, let $k = \\mathbb{R}, L = \\mathbb{C}$, and $\\Lambda = C_3$, where the Galois group $G = C_2$ acts via inversion. Then we have</p>\n\n<p>$$H \\otimes_{\\mathbb{R}} \\mathbb{C} \\cong \\mathbb{C}[C_3] \\cong \\mathbb{C}[x]/(x^3 - 1)$$</p>\n\n<p>where $x$ is grouplike but the action of complex conjugation sends $x$ to $x^{-1} = x^2$. The fixed point subalgebra $H$ has a basis given by the identity $1$ and the \"real and imaginary parts\"</p>\n\n<p>$$c = \\frac{x + x^{-1}}{2}$$\n$$s = \\frac{x - x^{-1}}{2i}$$</p>\n\n<p>(it helps to think of $x$ as secretly being $\\omega = e^{\\frac{2\\pi i}{3}}$ here) of $x$, with comultiplication given by the sine and cosine angle addition formulas</p>\n\n<p>$$\\Delta(c) = c \\otimes c - s \\otimes s$$\n$$\\Delta(s) = c \\otimes s + s \\otimes c.$$</p>\n\n<p>So neither $c$ nor $s$ is grouplike, and in fact $H$ has no nontrivial grouplike elements except the identity. In $H \\otimes_{\\mathbb{R}} \\mathbb{C}$ the nontrivial grouplike elements are $c + is$ and $c - is$. </p>\n" }, { "answer_id": 259613, "score": 6, "text": "<p>The case of irreducible, cocommutative Hopf algebras, over a field with $char(k)&gt; 0$, is discussed in Sweedler's textbook on Hopf algebras, Ch.$XIII$, sect. $13.2$. (See prop. $13.2.2$, $13.2.3$). </p>\n\n<p>For the case of cocommutative Hopf algebras, over an algebraically closed field $k$, with $char(k)\\geq 3$ you can have a look at: </p>\n\n<p><a href=\"https://www.semanticscholar.org/paper/Modules-of-Solvable-Innnitesimal-Groups-and-the-Farnsteiner-Voigt/a7ffd5538cd3e73320773c05ba4311e12e12905d\" rel=\"nofollow noreferrer\">Modules of solvable infinitesimal groups and the structure of representation-finite cocommutative Hopf algebras</a>, R.Farnsteiner, D.Voigt, Math. Proc. Cambridge Philos. Soc., v.127, p.441-459, 1999 </p>\n\n<p>and the references therein (among them, there is also an interesting older paper by the same authors discussing the case of cocommutative hopf algebras of finite representation type, over an algebraically closed field of $char(k)&gt;0$). </p>\n\n<p><strong>Edit (July 2018):</strong> Since user's Qiaochu Yuan answer, provides a form of counter-example to the theorem, for non-algebraically closed fields, i think it would be of some added value to indicate cases in which the Cartier-Kostant-Milnor-Moore theorem remains valid for non-algebraically closed fields: this happens if we take as an assumption that we are dealing with <a href=\"http://library.msri.org/books/Book43/files/andrus.pdf\" rel=\"nofollow noreferrer\">pointed</a> hopf algebras. So, the (somewhat more general) statement of the theorem goes like: </p>\n\n<blockquote>\n <p>If $H$ is a pointed, cocommutative hopf algebra over a field $k$ of characteristic $0$, we have that \n $$\nH\\cong U\\big(P(H)\\big)\\ltimes kG(H)\n$$\n where $P(H)$ is the lie algebra of the primitives, $U(.)$ its UEA and $G(H)$ the group of the group-likes. </p>\n</blockquote>\n\n<p>The algebraically closed field case (mentioned in the OP), may be considered a special case of the above, taking into account that cocommutative hopf algebras over algebraically closed fields, can be easily seen to be pointed. </p>\n" } ]
Question: Cartier-Kostant-Milnor-Moore theorem Answer 1: <p>When $k$ fails to be algebraically closed the theorem is false but the discrepancy can be understood in terms of Galois descent and so in principle understood in terms of Galois cohomology. </p> <p>Suppose $H$ is a cocommutative Hopf algebra over $k$, not algebraically closed but characteristic $0$. Then the classification theorem applies to the base change $H \otimes_k \bar{k}$ of $H$ to the algebraic closure, which must therefore take the form of the semidirect product of a universal enveloping algebra and a group algebra. So $H$ is a $k$-form of such a thing, but need not be such a thing itself.</p> <p>For example, suppose $H$ is finite-dimensional, so $H \otimes_k \bar{k}$ is a group algebra $\bar{k}[\Lambda]$ of a finite group $\Lambda$. The most obvious $k$-form of $\bar{k}[\Lambda]$ is, of course, $k[\Lambda]$, but other $k$-forms are possible. In this setting $H$ necessarily splits over a finite Galois extension $L$ in the sense that we must already have</p> <p>$$H \otimes_k L \cong L[\Lambda].$$</p> <p>Now, the Galois group $G = \text{Gal}(L/k)$ acts on the LHS by Hopf algebra automorphisms (over $k$), so also acts on the RHS by Hopf algebra automorphisms (over $k$), and hence acts on the group of grouplike elements $\Lambda$ on the RHS. If we already had $H \cong k[\Lambda]$ then this action would be trivial, but in fact every possible action of $G$ on $\Lambda$ occurs, and conjugacy classes of such actions classify the possible choices of $H$. Given such an action $\rho : H \to \text{Aut}(\Lambda)$ we can recover $H$ as</p> <p>$$H \cong L[\Lambda]^G$$</p> <p>where the action of $G$ on $L[\Lambda]$ extends the action of $G$ on $\Lambda$.</p> <p>Very explicitly, let $k = \mathbb{R}, L = \mathbb{C}$, and $\Lambda = C_3$, where the Galois group $G = C_2$ acts via inversion. Then we have</p> <p>$$H \otimes_{\mathbb{R}} \mathbb{C} \cong \mathbb{C}[C_3] \cong \mathbb{C}[x]/(x^3 - 1)$$</p> <p>where $x$ is grouplike but the action of complex conjugation sends $x$ to $x^{-1} = x^2$. The fixed point subalgebra $H$ has a basis given by the identity $1$ and the "real and imaginary parts"</p> <p>$$c = \frac{x + x^{-1}}{2}$$ $$s = \frac{x - x^{-1}}{2i}$$</p> <p>(it helps to think of $x$ as secretly being $\omega = e^{\frac{2\pi i}{3}}$ here) of $x$, with comultiplication given by the sine and cosine angle addition formulas</p> <p>$$\Delta(c) = c \otimes c - s \otimes s$$ $$\Delta(s) = c \otimes s + s \otimes c.$$</p> <p>So neither $c$ nor $s$ is grouplike, and in fact $H$ has no nontrivial grouplike elements except the identity. In $H \otimes_{\mathbb{R}} \mathbb{C}$ the nontrivial grouplike elements are $c + is$ and $c - is$. </p> Answer 2: <p>The case of irreducible, cocommutative Hopf algebras, over a field with $char(k)&gt; 0$, is discussed in Sweedler's textbook on Hopf algebras, Ch.$XIII$, sect. $13.2$. (See prop. $13.2.2$, $13.2.3$). </p> <p>For the case of cocommutative Hopf algebras, over an algebraically closed field $k$, with $char(k)\geq 3$ you can have a look at: </p> <p><a href="https://www.semanticscholar.org/paper/Modules-of-Solvable-Innnitesimal-Groups-and-the-Farnsteiner-Voigt/a7ffd5538cd3e73320773c05ba4311e12e12905d" rel="nofollow noreferrer">Modules of solvable infinitesimal groups and the structure of representation-finite cocommutative Hopf algebras</a>, R.Farnsteiner, D.Voigt, Math. Proc. Cambridge Philos. Soc., v.127, p.441-459, 1999 </p> <p>and the references therein (among them, there is also an interesting older paper by the same authors discussing the case of cocommutative hopf algebras of finite representation type, over an algebraically closed field of $char(k)&gt;0$). </p> <p><strong>Edit (July 2018):</strong> Since user's Qiaochu Yuan answer, provides a form of counter-example to the theorem, for non-algebraically closed fields, i think it would be of some added value to indicate cases in which the Cartier-Kostant-Milnor-Moore theorem remains valid for non-algebraically closed fields: this happens if we take as an assumption that we are dealing with <a href="http://library.msri.org/books/Book43/files/andrus.pdf" rel="nofollow noreferrer">pointed</a> hopf algebras. So, the (somewhat more general) statement of the theorem goes like: </p> <blockquote> <p>If $H$ is a pointed, cocommutative hopf algebra over a field $k$ of characteristic $0$, we have that $$ H\cong U\big(P(H)\big)\ltimes kG(H) $$ where $P(H)$ is the lie algebra of the primitives, $U(.)$ its UEA and $G(H)$ the group of the group-likes. </p> </blockquote> <p>The algebraically closed field case (mentioned in the OP), may be considered a special case of the above, taking into account that cocommutative hopf algebras over algebraically closed fields, can be easily seen to be pointed. </p> Answer 3: <p>When $k$ fails to be algebraically closed the theorem is false but the discrepancy can be understood in terms of Galois descent and so in principle understood in terms of Galois cohomology. </p> <p>Suppose $H$ is a cocommutative Hopf algebra over $k$, not algebraically closed but characteristic $0$. Then the classification theorem applies to the base change $H \otimes_k \bar{k}$ of $H$ to the algebraic closure, which must therefore take the form of the semidirect product of a universal enveloping algebra and a group algebra. So $H$ is a $k$-form of such a thing, but need not be such a thing itself.</p> <p>For example, suppose $H$ is finite-dimensional, so $H \otimes_k \bar{k}$ is a group algebra $\bar{k}[\Lambda]$ of a finite group $\Lambda$. The most obvious $k$-form of $\bar{k}[\Lambda]$ is, of course, $k[\Lambda]$, but other $k$-forms are possible. In this setting $H$ necessarily splits over a finite Galois extension $L$ in the sense that we must already have</p> <p>$$H \otimes_k L \cong L[\Lambda].$$</p> <p>Now, the Galois group $G = \text{Gal}(L/k)$ acts on the LHS by Hopf algebra automorphisms (over $k$), so also acts on the RHS by Hopf algebra automorphisms (over $k$), and hence acts on the group of grouplike elements $\Lambda$ on the RHS. If we already had $H \cong k[\Lambda]$ then this action would be trivial, but in fact every possible action of $G$ on $\Lambda$ occurs, and conjugacy classes of such actions classify the possible choices of $H$. Given such an action $\rho : H \to \text{Aut}(\Lambda)$ we can recover $H$ as</p> <p>$$H \cong L[\Lambda]^G$$</p> <p>where the action of $G$ on $L[\Lambda]$ extends the action of $G$ on $\Lambda$.</p> <p>Very explicitly, let $k = \mathbb{R}, L = \mathbb{C}$, and $\Lambda = C_3$, where the Galois group $G = C_2$ acts via inversion. Then we have</p> <p>$$H \otimes_{\mathbb{R}} \mathbb{C} \cong \mathbb{C}[C_3] \cong \mathbb{C}[x]/(x^3 - 1)$$</p> <p>where $x$ is grouplike but the action of complex conjugation sends $x$ to $x^{-1} = x^2$. The fixed point subalgebra $H$ has a basis given by the identity $1$ and the "real and imaginary parts"</p> <p>$$c = \frac{x + x^{-1}}{2}$$ $$s = \frac{x - x^{-1}}{2i}$$</p> <p>(it helps to think of $x$ as secretly being $\omega = e^{\frac{2\pi i}{3}}$ here) of $x$, with comultiplication given by the sine and cosine angle addition formulas</p> <p>$$\Delta(c) = c \otimes c - s \otimes s$$ $$\Delta(s) = c \otimes s + s \otimes c.$$</p> <p>So neither $c$ nor $s$ is grouplike, and in fact $H$ has no nontrivial grouplike elements except the identity. In $H \otimes_{\mathbb{R}} \mathbb{C}$ the nontrivial grouplike elements are $c + is$ and $c - is$. </p> Answer 4: <p>The case of irreducible, cocommutative Hopf algebras, over a field with $char(k)&gt; 0$, is discussed in Sweedler's textbook on Hopf algebras, Ch.$XIII$, sect. $13.2$. (See prop. $13.2.2$, $13.2.3$). </p> <p>For the case of cocommutative Hopf algebras, over an algebraically closed field $k$, with $char(k)\geq 3$ you can have a look at: </p> <p><a href="https://www.semanticscholar.org/paper/Modules-of-Solvable-Innnitesimal-Groups-and-the-Farnsteiner-Voigt/a7ffd5538cd3e73320773c05ba4311e12e12905d" rel="nofollow noreferrer">Modules of solvable infinitesimal groups and the structure of representation-finite cocommutative Hopf algebras</a>, R.Farnsteiner, D.Voigt, Math. Proc. Cambridge Philos. Soc., v.127, p.441-459, 1999 </p> <p>and the references therein (among them, there is also an interesting older paper by the same authors discussing the case of cocommutative hopf algebras of finite representation type, over an algebraically closed field of $char(k)&gt;0$). </p> <p><strong>Edit (July 2018):</strong> Since user's Qiaochu Yuan answer, provides a form of counter-example to the theorem, for non-algebraically closed fields, i think it would be of some added value to indicate cases in which the Cartier-Kostant-Milnor-Moore theorem remains valid for non-algebraically closed fields: this happens if we take as an assumption that we are dealing with <a href="http://library.msri.org/books/Book43/files/andrus.pdf" rel="nofollow noreferrer">pointed</a> hopf algebras. So, the (somewhat more general) statement of the theorem goes like: </p> <blockquote> <p>If $H$ is a pointed, cocommutative hopf algebra over a field $k$ of characteristic $0$, we have that $$ H\cong U\big(P(H)\big)\ltimes kG(H) $$ where $P(H)$ is the lie algebra of the primitives, $U(.)$ its UEA and $G(H)$ the group of the group-likes. </p> </blockquote> <p>The algebraically closed field case (mentioned in the OP), may be considered a special case of the above, taking into account that cocommutative hopf algebras over algebraically closed fields, can be easily seen to be pointed. </p> Answer 5: <p>When $k$ fails to be algebraically closed the theorem is false but the discrepancy can be understood in terms of Galois descent and so in principle understood in terms of Galois cohomology. </p> <p>Suppose $H$ is a cocommutative Hopf algebra over $k$, not algebraically closed but characteristic $0$. Then the classification theorem applies to the base change $H \otimes_k \bar{k}$ of $H$ to the algebraic closure, which must therefore take the form of the semidirect product of a universal enveloping algebra and a group algebra. So $H$ is a $k$-form of such a thing, but need not be such a thing itself.</p> <p>For example, suppose $H$ is finite-dimensional, so $H \otimes_k \bar{k}$ is a group algebra $\bar{k}[\Lambda]$ of a finite group $\Lambda$. The most obvious $k$-form of $\bar{k}[\Lambda]$ is, of course, $k[\Lambda]$, but other $k$-forms are possible. In this setting $H$ necessarily splits over a finite Galois extension $L$ in the sense that we must already have</p> <p>$$H \otimes_k L \cong L[\Lambda].$$</p> <p>Now, the Galois group $G = \text{Gal}(L/k)$ acts on the LHS by Hopf algebra automorphisms (over $k$), so also acts on the RHS by Hopf algebra automorphisms (over $k$), and hence acts on the group of grouplike elements $\Lambda$ on the RHS. If we already had $H \cong k[\Lambda]$ then this action would be trivial, but in fact every possible action of $G$ on $\Lambda$ occurs, and conjugacy classes of such actions classify the possible choices of $H$. Given such an action $\rho : H \to \text{Aut}(\Lambda)$ we can recover $H$ as</p> <p>$$H \cong L[\Lambda]^G$$</p> <p>where the action of $G$ on $L[\Lambda]$ extends the action of $G$ on $\Lambda$.</p> <p>Very explicitly, let $k = \mathbb{R}, L = \mathbb{C}$, and $\Lambda = C_3$, where the Galois group $G = C_2$ acts via inversion. Then we have</p> <p>$$H \otimes_{\mathbb{R}} \mathbb{C} \cong \mathbb{C}[C_3] \cong \mathbb{C}[x]/(x^3 - 1)$$</p> <p>where $x$ is grouplike but the action of complex conjugation sends $x$ to $x^{-1} = x^2$. The fixed point subalgebra $H$ has a basis given by the identity $1$ and the "real and imaginary parts"</p> <p>$$c = \frac{x + x^{-1}}{2}$$ $$s = \frac{x - x^{-1}}{2i}$$</p> <p>(it helps to think of $x$ as secretly being $\omega = e^{\frac{2\pi i}{3}}$ here) of $x$, with comultiplication given by the sine and cosine angle addition formulas</p> <p>$$\Delta(c) = c \otimes c - s \otimes s$$ $$\Delta(s) = c \otimes s + s \otimes c.$$</p> <p>So neither $c$ nor $s$ is grouplike, and in fact $H$ has no nontrivial grouplike elements except the identity. In $H \otimes_{\mathbb{R}} \mathbb{C}$ the nontrivial grouplike elements are $c + is$ and $c - is$. </p> Answer 6: <p>The case of irreducible, cocommutative Hopf algebras, over a field with $char(k)&gt; 0$, is discussed in Sweedler's textbook on Hopf algebras, Ch.$XIII$, sect. $13.2$. (See prop. $13.2.2$, $13.2.3$). </p> <p>For the case of cocommutative Hopf algebras, over an algebraically closed field $k$, with $char(k)\geq 3$ you can have a look at: </p> <p><a href="https://www.semanticscholar.org/paper/Modules-of-Solvable-Innnitesimal-Groups-and-the-Farnsteiner-Voigt/a7ffd5538cd3e73320773c05ba4311e12e12905d" rel="nofollow noreferrer">Modules of solvable infinitesimal groups and the structure of representation-finite cocommutative Hopf algebras</a>, R.Farnsteiner, D.Voigt, Math. Proc. Cambridge Philos. Soc., v.127, p.441-459, 1999 </p> <p>and the references therein (among them, there is also an interesting older paper by the same authors discussing the case of cocommutative hopf algebras of finite representation type, over an algebraically closed field of $char(k)&gt;0$). </p> <p><strong>Edit (July 2018):</strong> Since user's Qiaochu Yuan answer, provides a form of counter-example to the theorem, for non-algebraically closed fields, i think it would be of some added value to indicate cases in which the Cartier-Kostant-Milnor-Moore theorem remains valid for non-algebraically closed fields: this happens if we take as an assumption that we are dealing with <a href="http://library.msri.org/books/Book43/files/andrus.pdf" rel="nofollow noreferrer">pointed</a> hopf algebras. So, the (somewhat more general) statement of the theorem goes like: </p> <blockquote> <p>If $H$ is a pointed, cocommutative hopf algebra over a field $k$ of characteristic $0$, we have that $$ H\cong U\big(P(H)\big)\ltimes kG(H) $$ where $P(H)$ is the lie algebra of the primitives, $U(.)$ its UEA and $G(H)$ the group of the group-likes. </p> </blockquote> <p>The algebraically closed field case (mentioned in the OP), may be considered a special case of the above, taking into account that cocommutative hopf algebras over algebraically closed fields, can be easily seen to be pointed. </p>
Notations for dual spaces and dual operators
259,646
11
https://mathoverflow.net/questions/259646/notations-for-dual-spaces-and-dual-operators
<p>I'm asking for opinions about the 'best' notations for: 1. the algebraic dual of a vector space $X$; 2. the continuous dual of a TVS; 3. the algebraic dual (transpose) of an operator $T$ between vector spaces; 4. the dual (transpose) of a continuous operator between TVS; 5. the adjoint of a bounded operator $T$ between Hilbert spaces.</p> <p>My problem is that I would like to use these notions in the same context. The standard notations tend to overlap but I am forced to use different notations for each of these entities. Of course it is easy to come up with notations, but some traditions are well established and it is not trivial to respect them and at the same time keep them apart, with some elegance.</p> <p>What I'm using now: 1. $X'_{alg}$ 2. $X'$ 3. $^tT$ 4. $T'$ 5. $T^*$</p> <p>Thank you for your advice.</p>
[ { "answer_id": 259649, "score": 5, "text": "<p>The standard notation don't overlap as much as you think if you are careful with types. For example $X'$ is perfectly fine for both 1. and 2. because vector spaces and topological vector spaces are not the same type. Any object $X$ should only have one type so that there can never be any confusion. You can use the same notation if you never forget forgetful functors (i.e. never pretend that a TVS is the same as its underlying vector space). Similarly ' can be used for both 3. and 4.</p>\n\n<p>If you do want to extract the underlying vector space of a TVS some times, you might have an easier time just giving a name to the forgetful functor instead of inventing more notation.</p>\n\n<p>EDIT:</p>\n\n<p>And by the way: You don't <em>have</em> to name the forgetful functor $F$ or something similar like Nate used in the comment. You can also name it implicitely, although that solves only half of your problems. Simply stop using the abuse-of-notation $X$ when you mean $(X,\\tau)$ for a vector space topology $\\tau$ (which we already knew could happen when we introduced this and any other abuse-of-notation. They live only as long as they're useful and don't lead to confusion!) and the functor becomes $(X,\\tau)\\mapsto X$. Then you can use both $(X,\\tau)'$ and $X'$ without confusion. Of course problems 3. and 4. persist.</p>\n" }, { "answer_id": 271276, "score": 5, "text": "<p>Given your situation of having to juggle all these notational traditions at the same time, I would recommend for a space $X$ and an operator $A$:</p>\n\n<ol>\n<li>$X^{\\vee}$, 2. $X'$, 3. $A^{\\rm T}$, 4. $A^{\\rm T}$, 5. $A^*$</li>\n</ol>\n\n<p>My rationale is as follows. </p>\n\n<p>For 1: <strong>Algebraic</strong> geometers especially use the \"vee' notation for the algebraic dual so I think the cultural association helps the brain automatically make the association with the algebraic notion of dual.</p>\n\n<p>For 2: One usually does not write the space of tempered distributions as $S^*(\\mathbb{R}^d)$ but rather as $S'(\\mathbb{R^d})$. </p>\n\n<p>For 3 and 4: I would use the same notation as per the comment by user95282 since the maps are related by restriction. One could use ${}^{\\rm t}A$ for both but I prefer the matrix algebra notation if only because it is easier to type.</p>\n\n<p>For 5: As yuggib said, it is standard in the theory of $C^*$-algebras and spectral theory. I don't see any reason to be a contrarian and not follow what everyone else does.</p>\n" }, { "answer_id": 259668, "score": 2, "text": "<p>I like the notation $T^*$ for the adjoint, since it agrees with the usual notation for the involution of *-algebras. I don't know however if I would distinguish between $^t T$ and its continuous version with an <em>ad hoc</em> notation.</p>\n\n<p>For algebraic and continuous duals of (topological) vector spaces, personally I like Bourbaki's notation $X^*$ for the algebraic dual and $X'$ for the continuous dual (even if $X_{\\prime}$ for preduals is afwul, as opposed to $X_*$ that is natural if $X^*$ is the continuous dual). In addition, if $X$ is also a (subset of some) *-algebra, the notation $X^*$ may yield some confusion.</p>\n" }, { "answer_id": 259649, "score": 5, "text": "<p>The standard notation don't overlap as much as you think if you are careful with types. For example $X'$ is perfectly fine for both 1. and 2. because vector spaces and topological vector spaces are not the same type. Any object $X$ should only have one type so that there can never be any confusion. You can use the same notation if you never forget forgetful functors (i.e. never pretend that a TVS is the same as its underlying vector space). Similarly ' can be used for both 3. and 4.</p>\n\n<p>If you do want to extract the underlying vector space of a TVS some times, you might have an easier time just giving a name to the forgetful functor instead of inventing more notation.</p>\n\n<p>EDIT:</p>\n\n<p>And by the way: You don't <em>have</em> to name the forgetful functor $F$ or something similar like Nate used in the comment. You can also name it implicitely, although that solves only half of your problems. Simply stop using the abuse-of-notation $X$ when you mean $(X,\\tau)$ for a vector space topology $\\tau$ (which we already knew could happen when we introduced this and any other abuse-of-notation. They live only as long as they're useful and don't lead to confusion!) and the functor becomes $(X,\\tau)\\mapsto X$. Then you can use both $(X,\\tau)'$ and $X'$ without confusion. Of course problems 3. and 4. persist.</p>\n" }, { "answer_id": 271276, "score": 5, "text": "<p>Given your situation of having to juggle all these notational traditions at the same time, I would recommend for a space $X$ and an operator $A$:</p>\n\n<ol>\n<li>$X^{\\vee}$, 2. $X'$, 3. $A^{\\rm T}$, 4. $A^{\\rm T}$, 5. $A^*$</li>\n</ol>\n\n<p>My rationale is as follows. </p>\n\n<p>For 1: <strong>Algebraic</strong> geometers especially use the \"vee' notation for the algebraic dual so I think the cultural association helps the brain automatically make the association with the algebraic notion of dual.</p>\n\n<p>For 2: One usually does not write the space of tempered distributions as $S^*(\\mathbb{R}^d)$ but rather as $S'(\\mathbb{R^d})$. </p>\n\n<p>For 3 and 4: I would use the same notation as per the comment by user95282 since the maps are related by restriction. One could use ${}^{\\rm t}A$ for both but I prefer the matrix algebra notation if only because it is easier to type.</p>\n\n<p>For 5: As yuggib said, it is standard in the theory of $C^*$-algebras and spectral theory. I don't see any reason to be a contrarian and not follow what everyone else does.</p>\n" }, { "answer_id": 259668, "score": 2, "text": "<p>I like the notation $T^*$ for the adjoint, since it agrees with the usual notation for the involution of *-algebras. I don't know however if I would distinguish between $^t T$ and its continuous version with an <em>ad hoc</em> notation.</p>\n\n<p>For algebraic and continuous duals of (topological) vector spaces, personally I like Bourbaki's notation $X^*$ for the algebraic dual and $X'$ for the continuous dual (even if $X_{\\prime}$ for preduals is afwul, as opposed to $X_*$ that is natural if $X^*$ is the continuous dual). In addition, if $X$ is also a (subset of some) *-algebra, the notation $X^*$ may yield some confusion.</p>\n" }, { "answer_id": 259649, "score": 5, "text": "<p>The standard notation don't overlap as much as you think if you are careful with types. For example $X'$ is perfectly fine for both 1. and 2. because vector spaces and topological vector spaces are not the same type. Any object $X$ should only have one type so that there can never be any confusion. You can use the same notation if you never forget forgetful functors (i.e. never pretend that a TVS is the same as its underlying vector space). Similarly ' can be used for both 3. and 4.</p>\n\n<p>If you do want to extract the underlying vector space of a TVS some times, you might have an easier time just giving a name to the forgetful functor instead of inventing more notation.</p>\n\n<p>EDIT:</p>\n\n<p>And by the way: You don't <em>have</em> to name the forgetful functor $F$ or something similar like Nate used in the comment. You can also name it implicitely, although that solves only half of your problems. Simply stop using the abuse-of-notation $X$ when you mean $(X,\\tau)$ for a vector space topology $\\tau$ (which we already knew could happen when we introduced this and any other abuse-of-notation. They live only as long as they're useful and don't lead to confusion!) and the functor becomes $(X,\\tau)\\mapsto X$. Then you can use both $(X,\\tau)'$ and $X'$ without confusion. Of course problems 3. and 4. persist.</p>\n" }, { "answer_id": 271276, "score": 5, "text": "<p>Given your situation of having to juggle all these notational traditions at the same time, I would recommend for a space $X$ and an operator $A$:</p>\n\n<ol>\n<li>$X^{\\vee}$, 2. $X'$, 3. $A^{\\rm T}$, 4. $A^{\\rm T}$, 5. $A^*$</li>\n</ol>\n\n<p>My rationale is as follows. </p>\n\n<p>For 1: <strong>Algebraic</strong> geometers especially use the \"vee' notation for the algebraic dual so I think the cultural association helps the brain automatically make the association with the algebraic notion of dual.</p>\n\n<p>For 2: One usually does not write the space of tempered distributions as $S^*(\\mathbb{R}^d)$ but rather as $S'(\\mathbb{R^d})$. </p>\n\n<p>For 3 and 4: I would use the same notation as per the comment by user95282 since the maps are related by restriction. One could use ${}^{\\rm t}A$ for both but I prefer the matrix algebra notation if only because it is easier to type.</p>\n\n<p>For 5: As yuggib said, it is standard in the theory of $C^*$-algebras and spectral theory. I don't see any reason to be a contrarian and not follow what everyone else does.</p>\n" }, { "answer_id": 259668, "score": 2, "text": "<p>I like the notation $T^*$ for the adjoint, since it agrees with the usual notation for the involution of *-algebras. I don't know however if I would distinguish between $^t T$ and its continuous version with an <em>ad hoc</em> notation.</p>\n\n<p>For algebraic and continuous duals of (topological) vector spaces, personally I like Bourbaki's notation $X^*$ for the algebraic dual and $X'$ for the continuous dual (even if $X_{\\prime}$ for preduals is afwul, as opposed to $X_*$ that is natural if $X^*$ is the continuous dual). In addition, if $X$ is also a (subset of some) *-algebra, the notation $X^*$ may yield some confusion.</p>\n" } ]
Question: Notations for dual spaces and dual operators Answer 1: <p>The standard notation don't overlap as much as you think if you are careful with types. For example $X'$ is perfectly fine for both 1. and 2. because vector spaces and topological vector spaces are not the same type. Any object $X$ should only have one type so that there can never be any confusion. You can use the same notation if you never forget forgetful functors (i.e. never pretend that a TVS is the same as its underlying vector space). Similarly ' can be used for both 3. and 4.</p> <p>If you do want to extract the underlying vector space of a TVS some times, you might have an easier time just giving a name to the forgetful functor instead of inventing more notation.</p> <p>EDIT:</p> <p>And by the way: You don't <em>have</em> to name the forgetful functor $F$ or something similar like Nate used in the comment. You can also name it implicitely, although that solves only half of your problems. Simply stop using the abuse-of-notation $X$ when you mean $(X,\tau)$ for a vector space topology $\tau$ (which we already knew could happen when we introduced this and any other abuse-of-notation. They live only as long as they're useful and don't lead to confusion!) and the functor becomes $(X,\tau)\mapsto X$. Then you can use both $(X,\tau)'$ and $X'$ without confusion. Of course problems 3. and 4. persist.</p> Answer 2: <p>Given your situation of having to juggle all these notational traditions at the same time, I would recommend for a space $X$ and an operator $A$:</p> <ol> <li>$X^{\vee}$, 2. $X'$, 3. $A^{\rm T}$, 4. $A^{\rm T}$, 5. $A^*$</li> </ol> <p>My rationale is as follows. </p> <p>For 1: <strong>Algebraic</strong> geometers especially use the "vee' notation for the algebraic dual so I think the cultural association helps the brain automatically make the association with the algebraic notion of dual.</p> <p>For 2: One usually does not write the space of tempered distributions as $S^*(\mathbb{R}^d)$ but rather as $S'(\mathbb{R^d})$. </p> <p>For 3 and 4: I would use the same notation as per the comment by user95282 since the maps are related by restriction. One could use ${}^{\rm t}A$ for both but I prefer the matrix algebra notation if only because it is easier to type.</p> <p>For 5: As yuggib said, it is standard in the theory of $C^*$-algebras and spectral theory. I don't see any reason to be a contrarian and not follow what everyone else does.</p> Answer 3: <p>I like the notation $T^*$ for the adjoint, since it agrees with the usual notation for the involution of *-algebras. I don't know however if I would distinguish between $^t T$ and its continuous version with an <em>ad hoc</em> notation.</p> <p>For algebraic and continuous duals of (topological) vector spaces, personally I like Bourbaki's notation $X^*$ for the algebraic dual and $X'$ for the continuous dual (even if $X_{\prime}$ for preduals is afwul, as opposed to $X_*$ that is natural if $X^*$ is the continuous dual). In addition, if $X$ is also a (subset of some) *-algebra, the notation $X^*$ may yield some confusion.</p> Answer 4: <p>The standard notation don't overlap as much as you think if you are careful with types. For example $X'$ is perfectly fine for both 1. and 2. because vector spaces and topological vector spaces are not the same type. Any object $X$ should only have one type so that there can never be any confusion. You can use the same notation if you never forget forgetful functors (i.e. never pretend that a TVS is the same as its underlying vector space). Similarly ' can be used for both 3. and 4.</p> <p>If you do want to extract the underlying vector space of a TVS some times, you might have an easier time just giving a name to the forgetful functor instead of inventing more notation.</p> <p>EDIT:</p> <p>And by the way: You don't <em>have</em> to name the forgetful functor $F$ or something similar like Nate used in the comment. You can also name it implicitely, although that solves only half of your problems. Simply stop using the abuse-of-notation $X$ when you mean $(X,\tau)$ for a vector space topology $\tau$ (which we already knew could happen when we introduced this and any other abuse-of-notation. They live only as long as they're useful and don't lead to confusion!) and the functor becomes $(X,\tau)\mapsto X$. Then you can use both $(X,\tau)'$ and $X'$ without confusion. Of course problems 3. and 4. persist.</p> Answer 5: <p>Given your situation of having to juggle all these notational traditions at the same time, I would recommend for a space $X$ and an operator $A$:</p> <ol> <li>$X^{\vee}$, 2. $X'$, 3. $A^{\rm T}$, 4. $A^{\rm T}$, 5. $A^*$</li> </ol> <p>My rationale is as follows. </p> <p>For 1: <strong>Algebraic</strong> geometers especially use the "vee' notation for the algebraic dual so I think the cultural association helps the brain automatically make the association with the algebraic notion of dual.</p> <p>For 2: One usually does not write the space of tempered distributions as $S^*(\mathbb{R}^d)$ but rather as $S'(\mathbb{R^d})$. </p> <p>For 3 and 4: I would use the same notation as per the comment by user95282 since the maps are related by restriction. One could use ${}^{\rm t}A$ for both but I prefer the matrix algebra notation if only because it is easier to type.</p> <p>For 5: As yuggib said, it is standard in the theory of $C^*$-algebras and spectral theory. I don't see any reason to be a contrarian and not follow what everyone else does.</p> Answer 6: <p>I like the notation $T^*$ for the adjoint, since it agrees with the usual notation for the involution of *-algebras. I don't know however if I would distinguish between $^t T$ and its continuous version with an <em>ad hoc</em> notation.</p> <p>For algebraic and continuous duals of (topological) vector spaces, personally I like Bourbaki's notation $X^*$ for the algebraic dual and $X'$ for the continuous dual (even if $X_{\prime}$ for preduals is afwul, as opposed to $X_*$ that is natural if $X^*$ is the continuous dual). In addition, if $X$ is also a (subset of some) *-algebra, the notation $X^*$ may yield some confusion.</p> Answer 7: <p>The standard notation don't overlap as much as you think if you are careful with types. For example $X'$ is perfectly fine for both 1. and 2. because vector spaces and topological vector spaces are not the same type. Any object $X$ should only have one type so that there can never be any confusion. You can use the same notation if you never forget forgetful functors (i.e. never pretend that a TVS is the same as its underlying vector space). Similarly ' can be used for both 3. and 4.</p> <p>If you do want to extract the underlying vector space of a TVS some times, you might have an easier time just giving a name to the forgetful functor instead of inventing more notation.</p> <p>EDIT:</p> <p>And by the way: You don't <em>have</em> to name the forgetful functor $F$ or something similar like Nate used in the comment. You can also name it implicitely, although that solves only half of your problems. Simply stop using the abuse-of-notation $X$ when you mean $(X,\tau)$ for a vector space topology $\tau$ (which we already knew could happen when we introduced this and any other abuse-of-notation. They live only as long as they're useful and don't lead to confusion!) and the functor becomes $(X,\tau)\mapsto X$. Then you can use both $(X,\tau)'$ and $X'$ without confusion. Of course problems 3. and 4. persist.</p> Answer 8: <p>Given your situation of having to juggle all these notational traditions at the same time, I would recommend for a space $X$ and an operator $A$:</p> <ol> <li>$X^{\vee}$, 2. $X'$, 3. $A^{\rm T}$, 4. $A^{\rm T}$, 5. $A^*$</li> </ol> <p>My rationale is as follows. </p> <p>For 1: <strong>Algebraic</strong> geometers especially use the "vee' notation for the algebraic dual so I think the cultural association helps the brain automatically make the association with the algebraic notion of dual.</p> <p>For 2: One usually does not write the space of tempered distributions as $S^*(\mathbb{R}^d)$ but rather as $S'(\mathbb{R^d})$. </p> <p>For 3 and 4: I would use the same notation as per the comment by user95282 since the maps are related by restriction. One could use ${}^{\rm t}A$ for both but I prefer the matrix algebra notation if only because it is easier to type.</p> <p>For 5: As yuggib said, it is standard in the theory of $C^*$-algebras and spectral theory. I don't see any reason to be a contrarian and not follow what everyone else does.</p> Answer 9: <p>I like the notation $T^*$ for the adjoint, since it agrees with the usual notation for the involution of *-algebras. I don't know however if I would distinguish between $^t T$ and its continuous version with an <em>ad hoc</em> notation.</p> <p>For algebraic and continuous duals of (topological) vector spaces, personally I like Bourbaki's notation $X^*$ for the algebraic dual and $X'$ for the continuous dual (even if $X_{\prime}$ for preduals is afwul, as opposed to $X_*$ that is natural if $X^*$ is the continuous dual). In addition, if $X$ is also a (subset of some) *-algebra, the notation $X^*$ may yield some confusion.</p>
What is a finite Haken cover of the Seifert–Weber space?
259,951
11
https://mathoverflow.net/questions/259951/what-is-a-finite-haken-cover-of-the-seifert-weber-space
<p>It's known that the Seifert–Weber space (obtained from a dodecahedron by gluing opposite faces with a 3/10 turn) is an example of a non-Haken 3-manifold. Since every closed 3-manifold is virtually Haken, I was wondering: is there was a known finite cover of the Seifert-Weber space that is Haken?</p>
[ { "answer_id": 259969, "score": 9, "text": "<p>I just checked using SnapPea that there is a cover of the Seifert-Weber dodecahedral space of index 25 (a 5-fold cyclic cover of a 5-fold cyclic cover) which has positive first betti number hence is Haken. </p>\n\n<p>The Seifert-Weber space is a 5-fold cyclic branched cover over the Whitehead link complement. There are two such 5-fold covers (up to homeomorphism), which one may compute using SnapPea (perform $(5,0)$ surgery on each cusp of the Whitehead link, then compute all 5-fold cyclic covers of this orbifold, giving four manifold covers, with two isometry types). One may then compute the 5-fold cyclic covers of these two manifolds. One of them (not Seifert-Weber) has a 5-fold cyclic cover with positive betti number, whereas the 5-fold cyclic covers of the Seifert-Weber space have trivial betti number. However, one of them will be a 5-fold cover of its sibling, and hence will have a 5-fold cyclic cover which has positive betti number. </p>\n\n<p>There are many other ways that we know the Seifert-Weber space to have a finite cover with positive first betti number (Hempel's paper pointed out by Igor shows that there is a 5-fold irregular cover), but it is an interesting question whether given a manifold $M$ with $b_1(M,\\mathbb{F}_p)\\geq 4$, is there a $p$-cover which has positive first betti number (I asked this as question 5 in a <a href=\"https://math.berkeley.edu/sites/default/files/faculty/files/virtualspecialICM.pdf\" rel=\"nofollow noreferrer\">survey paper)</a>. Since all 5-fold covers of the Seifert-Weber space have $b_1(*; \\mathbb{F}_5)\\geq 4$, then this computation shows that a single 5-fold cyclic cover works in some cases. </p>\n" }, { "answer_id": 259958, "score": 8, "text": "<p>This is constructed (reasonably explicitly) in John Hempel's 1982 paper.</p>\n\n<p><em>John Hempel</em>, MR 664329 <a href=\"http://dx.doi.org/10.1007/BF01393377\"><strong>Orientation reversing involutions and the first Betti number for finite coverings of $3$-manifolds</strong></a>, <em>Invent. Math.</em> <strong>67</strong> (1982), no. 1, 133--142.</p>\n" }, { "answer_id": 259969, "score": 9, "text": "<p>I just checked using SnapPea that there is a cover of the Seifert-Weber dodecahedral space of index 25 (a 5-fold cyclic cover of a 5-fold cyclic cover) which has positive first betti number hence is Haken. </p>\n\n<p>The Seifert-Weber space is a 5-fold cyclic branched cover over the Whitehead link complement. There are two such 5-fold covers (up to homeomorphism), which one may compute using SnapPea (perform $(5,0)$ surgery on each cusp of the Whitehead link, then compute all 5-fold cyclic covers of this orbifold, giving four manifold covers, with two isometry types). One may then compute the 5-fold cyclic covers of these two manifolds. One of them (not Seifert-Weber) has a 5-fold cyclic cover with positive betti number, whereas the 5-fold cyclic covers of the Seifert-Weber space have trivial betti number. However, one of them will be a 5-fold cover of its sibling, and hence will have a 5-fold cyclic cover which has positive betti number. </p>\n\n<p>There are many other ways that we know the Seifert-Weber space to have a finite cover with positive first betti number (Hempel's paper pointed out by Igor shows that there is a 5-fold irregular cover), but it is an interesting question whether given a manifold $M$ with $b_1(M,\\mathbb{F}_p)\\geq 4$, is there a $p$-cover which has positive first betti number (I asked this as question 5 in a <a href=\"https://math.berkeley.edu/sites/default/files/faculty/files/virtualspecialICM.pdf\" rel=\"nofollow noreferrer\">survey paper)</a>. Since all 5-fold covers of the Seifert-Weber space have $b_1(*; \\mathbb{F}_5)\\geq 4$, then this computation shows that a single 5-fold cyclic cover works in some cases. </p>\n" }, { "answer_id": 259958, "score": 8, "text": "<p>This is constructed (reasonably explicitly) in John Hempel's 1982 paper.</p>\n\n<p><em>John Hempel</em>, MR 664329 <a href=\"http://dx.doi.org/10.1007/BF01393377\"><strong>Orientation reversing involutions and the first Betti number for finite coverings of $3$-manifolds</strong></a>, <em>Invent. Math.</em> <strong>67</strong> (1982), no. 1, 133--142.</p>\n" }, { "answer_id": 259969, "score": 9, "text": "<p>I just checked using SnapPea that there is a cover of the Seifert-Weber dodecahedral space of index 25 (a 5-fold cyclic cover of a 5-fold cyclic cover) which has positive first betti number hence is Haken. </p>\n\n<p>The Seifert-Weber space is a 5-fold cyclic branched cover over the Whitehead link complement. There are two such 5-fold covers (up to homeomorphism), which one may compute using SnapPea (perform $(5,0)$ surgery on each cusp of the Whitehead link, then compute all 5-fold cyclic covers of this orbifold, giving four manifold covers, with two isometry types). One may then compute the 5-fold cyclic covers of these two manifolds. One of them (not Seifert-Weber) has a 5-fold cyclic cover with positive betti number, whereas the 5-fold cyclic covers of the Seifert-Weber space have trivial betti number. However, one of them will be a 5-fold cover of its sibling, and hence will have a 5-fold cyclic cover which has positive betti number. </p>\n\n<p>There are many other ways that we know the Seifert-Weber space to have a finite cover with positive first betti number (Hempel's paper pointed out by Igor shows that there is a 5-fold irregular cover), but it is an interesting question whether given a manifold $M$ with $b_1(M,\\mathbb{F}_p)\\geq 4$, is there a $p$-cover which has positive first betti number (I asked this as question 5 in a <a href=\"https://math.berkeley.edu/sites/default/files/faculty/files/virtualspecialICM.pdf\" rel=\"nofollow noreferrer\">survey paper)</a>. Since all 5-fold covers of the Seifert-Weber space have $b_1(*; \\mathbb{F}_5)\\geq 4$, then this computation shows that a single 5-fold cyclic cover works in some cases. </p>\n" }, { "answer_id": 259958, "score": 8, "text": "<p>This is constructed (reasonably explicitly) in John Hempel's 1982 paper.</p>\n\n<p><em>John Hempel</em>, MR 664329 <a href=\"http://dx.doi.org/10.1007/BF01393377\"><strong>Orientation reversing involutions and the first Betti number for finite coverings of $3$-manifolds</strong></a>, <em>Invent. Math.</em> <strong>67</strong> (1982), no. 1, 133--142.</p>\n" } ]
Question: What is a finite Haken cover of the Seifert–Weber space? Answer 1: <p>I just checked using SnapPea that there is a cover of the Seifert-Weber dodecahedral space of index 25 (a 5-fold cyclic cover of a 5-fold cyclic cover) which has positive first betti number hence is Haken. </p> <p>The Seifert-Weber space is a 5-fold cyclic branched cover over the Whitehead link complement. There are two such 5-fold covers (up to homeomorphism), which one may compute using SnapPea (perform $(5,0)$ surgery on each cusp of the Whitehead link, then compute all 5-fold cyclic covers of this orbifold, giving four manifold covers, with two isometry types). One may then compute the 5-fold cyclic covers of these two manifolds. One of them (not Seifert-Weber) has a 5-fold cyclic cover with positive betti number, whereas the 5-fold cyclic covers of the Seifert-Weber space have trivial betti number. However, one of them will be a 5-fold cover of its sibling, and hence will have a 5-fold cyclic cover which has positive betti number. </p> <p>There are many other ways that we know the Seifert-Weber space to have a finite cover with positive first betti number (Hempel's paper pointed out by Igor shows that there is a 5-fold irregular cover), but it is an interesting question whether given a manifold $M$ with $b_1(M,\mathbb{F}_p)\geq 4$, is there a $p$-cover which has positive first betti number (I asked this as question 5 in a <a href="https://math.berkeley.edu/sites/default/files/faculty/files/virtualspecialICM.pdf" rel="nofollow noreferrer">survey paper)</a>. Since all 5-fold covers of the Seifert-Weber space have $b_1(*; \mathbb{F}_5)\geq 4$, then this computation shows that a single 5-fold cyclic cover works in some cases. </p> Answer 2: <p>This is constructed (reasonably explicitly) in John Hempel's 1982 paper.</p> <p><em>John Hempel</em>, MR 664329 <a href="http://dx.doi.org/10.1007/BF01393377"><strong>Orientation reversing involutions and the first Betti number for finite coverings of $3$-manifolds</strong></a>, <em>Invent. Math.</em> <strong>67</strong> (1982), no. 1, 133--142.</p> Answer 3: <p>I just checked using SnapPea that there is a cover of the Seifert-Weber dodecahedral space of index 25 (a 5-fold cyclic cover of a 5-fold cyclic cover) which has positive first betti number hence is Haken. </p> <p>The Seifert-Weber space is a 5-fold cyclic branched cover over the Whitehead link complement. There are two such 5-fold covers (up to homeomorphism), which one may compute using SnapPea (perform $(5,0)$ surgery on each cusp of the Whitehead link, then compute all 5-fold cyclic covers of this orbifold, giving four manifold covers, with two isometry types). One may then compute the 5-fold cyclic covers of these two manifolds. One of them (not Seifert-Weber) has a 5-fold cyclic cover with positive betti number, whereas the 5-fold cyclic covers of the Seifert-Weber space have trivial betti number. However, one of them will be a 5-fold cover of its sibling, and hence will have a 5-fold cyclic cover which has positive betti number. </p> <p>There are many other ways that we know the Seifert-Weber space to have a finite cover with positive first betti number (Hempel's paper pointed out by Igor shows that there is a 5-fold irregular cover), but it is an interesting question whether given a manifold $M$ with $b_1(M,\mathbb{F}_p)\geq 4$, is there a $p$-cover which has positive first betti number (I asked this as question 5 in a <a href="https://math.berkeley.edu/sites/default/files/faculty/files/virtualspecialICM.pdf" rel="nofollow noreferrer">survey paper)</a>. Since all 5-fold covers of the Seifert-Weber space have $b_1(*; \mathbb{F}_5)\geq 4$, then this computation shows that a single 5-fold cyclic cover works in some cases. </p> Answer 4: <p>This is constructed (reasonably explicitly) in John Hempel's 1982 paper.</p> <p><em>John Hempel</em>, MR 664329 <a href="http://dx.doi.org/10.1007/BF01393377"><strong>Orientation reversing involutions and the first Betti number for finite coverings of $3$-manifolds</strong></a>, <em>Invent. Math.</em> <strong>67</strong> (1982), no. 1, 133--142.</p> Answer 5: <p>I just checked using SnapPea that there is a cover of the Seifert-Weber dodecahedral space of index 25 (a 5-fold cyclic cover of a 5-fold cyclic cover) which has positive first betti number hence is Haken. </p> <p>The Seifert-Weber space is a 5-fold cyclic branched cover over the Whitehead link complement. There are two such 5-fold covers (up to homeomorphism), which one may compute using SnapPea (perform $(5,0)$ surgery on each cusp of the Whitehead link, then compute all 5-fold cyclic covers of this orbifold, giving four manifold covers, with two isometry types). One may then compute the 5-fold cyclic covers of these two manifolds. One of them (not Seifert-Weber) has a 5-fold cyclic cover with positive betti number, whereas the 5-fold cyclic covers of the Seifert-Weber space have trivial betti number. However, one of them will be a 5-fold cover of its sibling, and hence will have a 5-fold cyclic cover which has positive betti number. </p> <p>There are many other ways that we know the Seifert-Weber space to have a finite cover with positive first betti number (Hempel's paper pointed out by Igor shows that there is a 5-fold irregular cover), but it is an interesting question whether given a manifold $M$ with $b_1(M,\mathbb{F}_p)\geq 4$, is there a $p$-cover which has positive first betti number (I asked this as question 5 in a <a href="https://math.berkeley.edu/sites/default/files/faculty/files/virtualspecialICM.pdf" rel="nofollow noreferrer">survey paper)</a>. Since all 5-fold covers of the Seifert-Weber space have $b_1(*; \mathbb{F}_5)\geq 4$, then this computation shows that a single 5-fold cyclic cover works in some cases. </p> Answer 6: <p>This is constructed (reasonably explicitly) in John Hempel's 1982 paper.</p> <p><em>John Hempel</em>, MR 664329 <a href="http://dx.doi.org/10.1007/BF01393377"><strong>Orientation reversing involutions and the first Betti number for finite coverings of $3$-manifolds</strong></a>, <em>Invent. Math.</em> <strong>67</strong> (1982), no. 1, 133--142.</p>
Where is Seminaire Bourbaki on-line?
260,474
11
https://mathoverflow.net/questions/260474/where-is-seminaire-bourbaki-on-line
<p>Where is Seminaire Bourbaki on-line?</p> <p>The reason I ask is that for years it was complicated-enough already to find it in traditional libraries, as it would be catalogued according to its venue or ephemeral names of various aspects of it.</p> <p>With the advent of the internet, because most of the Sem. Bourb. archive was so old, there was no obvious, easy way to convert it to digital, and, obviously, scanning-in cannot be done overnight.</p> <p>EDIT: and, to make clear(er) what has been so obvious to me for decades that I'd failed to explain it in the question... for many decades, Sem. Bourb. consisted of not-too-long superb expositions by superb mathematicians of other superb mathematicians' important work. (Although at some point many of these transcripts were in English, not French), as I say to my students, the existence of decades of Sem. Bourb. transcripts is reason enough to learn how to read mathematical French.</p>
[ { "answer_id": 326494, "score": 10, "text": "<p>This is an extension on Paul's answer, but it would have been too extensive of an edit so I posted this separately.</p>\n\n<p>(I have corrected the link in Paul's post since the old one was no longer working)</p>\n\n<blockquote>\n <p>Articles from <strong>1948</strong> (creation) up to <strong>2006</strong> can be found at <a href=\"http://www.numdam.org/actas/SB/\" rel=\"nofollow noreferrer\">this</a> link in Paul's post</p>\n \n <p>Articles <em>after</em> 2006 can also be purchased from SMF (Société Mathématique de France) at the following links: <a href=\"http://smf4.emath.fr/Publications/Asterisque/2008/317/html\" rel=\"nofollow noreferrer\">2006-2007</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2009/326/html\" rel=\"nofollow noreferrer\">2007-2008</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2010/332/html\" rel=\"nofollow noreferrer\">2008-2009</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2011/339/html\" rel=\"nofollow noreferrer\">2009-2010</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2012/348/html\" rel=\"nofollow noreferrer\">2010-2011</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2013/352/html\" rel=\"nofollow noreferrer\">2011-2012</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2014/361/html\" rel=\"nofollow noreferrer\">2012-2013</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2015/367-368/html\" rel=\"nofollow noreferrer\">2013-2014</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2016/380/html\" rel=\"nofollow noreferrer\">2014-2015</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2017/390/html\" rel=\"nofollow noreferrer\">2015-2016</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2017/390/html\" rel=\"nofollow noreferrer\">2016-2017</a>, (articles after 2017 can only be accessed, at the time of posting this, by clicking \"PDF\" next to the bolded article title at the following links: <a href=\"http://www.bourbaki.ens.fr/seminaires/2018/\" rel=\"nofollow noreferrer\">2018</a>, <a href=\"http://www.bourbaki.ens.fr/seminaires/2019/\" rel=\"nofollow noreferrer\">2019</a>).</p>\n</blockquote>\n\n<p><s>Note: Since the 2019 seminar is 3 days from the time of writing this, those files are not available at this time.</s></p>\n\n<p>As of April 1st, the articles from the March 2019 seminar are available at the link above.</p>\n\n<blockquote>\n <p>If you are wanting to download an article published <strong>after 2006</strong>, you just need to know the corresponding Exposé number which can be found under \"CONTENU DU VOLUME\" on the SMF links above, after which the article can be found at links with this format:\n <code>http://www.bourbaki.ens.fr/TEXTES/(put the exposé # here).pdf</code>. <br/>For example, <a href=\"http://www.bourbaki.ens.fr/TEXTES/1108.pdf\" rel=\"nofollow noreferrer\">http://www.bourbaki.ens.fr/TEXTES/1108.pdf</a></p>\n</blockquote>\n" }, { "answer_id": 260475, "score": 7, "text": "<p>Thanks to @nfdc's comment following up on a comment of mine to the question <a href=\"https://mathoverflow.net/questions/260407/compactness-of-the-automorphic-quotient/260426\">Compactness of the automorphic quotient</a>, I/we should know that Sem. Bourb. is archived in the numdam.org at</p>\n\n<p><a href=\"http://www.numdam.org/actas/SB/\" rel=\"nofollow noreferrer\">http://www.numdam.org/actas/SB/</a></p>\n\n<p>EDIT: As @მამუკაჯიბლაძე observes in a comment, this includes the seminar up through 2006.</p>\n" }, { "answer_id": 393457, "score": 5, "text": "<p>Here you have a complete and updated set of resources\n<a href=\"https://www.bourbaki.fr/index-en.html\" rel=\"noreferrer\">https://www.bourbaki.fr/index-en.html</a></p>\n<ul>\n<li> Elements of mathematics <br>\nand, in particular, the two new volumes (Topologie Algébrique and Théories Spectrales) \n<li> Séminaire Betty B. <br>\nThe junior Bourbaki Seminar\n<li> N. Bourbaki Seminar <br>\nThe very one <br>\nPrevious sessions \n<li> Recorded talks of the séminaire N. Bourbaki\n(on youtube)\n<li> Earlier Texts <br>\nVolumes of the Seminar since its creation. <br>\nTexts of the Bourbaki Seminar lectures until 2010 (on Numdam) \n<li> Poincaré Seminar <br>\nIf you like Physics.\n</ul>\n" }, { "answer_id": 326938, "score": 1, "text": "<p>If you are interested in the actual talks, rather than the books, you can find it on youtube, <a href=\"https://www.youtube.com/user/PoincareInstitute/videos\" rel=\"nofollow noreferrer\">here</a>. They also stream them in direct, during the talks !</p>\n" }, { "answer_id": 326494, "score": 10, "text": "<p>This is an extension on Paul's answer, but it would have been too extensive of an edit so I posted this separately.</p>\n\n<p>(I have corrected the link in Paul's post since the old one was no longer working)</p>\n\n<blockquote>\n <p>Articles from <strong>1948</strong> (creation) up to <strong>2006</strong> can be found at <a href=\"http://www.numdam.org/actas/SB/\" rel=\"nofollow noreferrer\">this</a> link in Paul's post</p>\n \n <p>Articles <em>after</em> 2006 can also be purchased from SMF (Société Mathématique de France) at the following links: <a href=\"http://smf4.emath.fr/Publications/Asterisque/2008/317/html\" rel=\"nofollow noreferrer\">2006-2007</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2009/326/html\" rel=\"nofollow noreferrer\">2007-2008</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2010/332/html\" rel=\"nofollow noreferrer\">2008-2009</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2011/339/html\" rel=\"nofollow noreferrer\">2009-2010</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2012/348/html\" rel=\"nofollow noreferrer\">2010-2011</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2013/352/html\" rel=\"nofollow noreferrer\">2011-2012</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2014/361/html\" rel=\"nofollow noreferrer\">2012-2013</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2015/367-368/html\" rel=\"nofollow noreferrer\">2013-2014</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2016/380/html\" rel=\"nofollow noreferrer\">2014-2015</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2017/390/html\" rel=\"nofollow noreferrer\">2015-2016</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2017/390/html\" rel=\"nofollow noreferrer\">2016-2017</a>, (articles after 2017 can only be accessed, at the time of posting this, by clicking \"PDF\" next to the bolded article title at the following links: <a href=\"http://www.bourbaki.ens.fr/seminaires/2018/\" rel=\"nofollow noreferrer\">2018</a>, <a href=\"http://www.bourbaki.ens.fr/seminaires/2019/\" rel=\"nofollow noreferrer\">2019</a>).</p>\n</blockquote>\n\n<p><s>Note: Since the 2019 seminar is 3 days from the time of writing this, those files are not available at this time.</s></p>\n\n<p>As of April 1st, the articles from the March 2019 seminar are available at the link above.</p>\n\n<blockquote>\n <p>If you are wanting to download an article published <strong>after 2006</strong>, you just need to know the corresponding Exposé number which can be found under \"CONTENU DU VOLUME\" on the SMF links above, after which the article can be found at links with this format:\n <code>http://www.bourbaki.ens.fr/TEXTES/(put the exposé # here).pdf</code>. <br/>For example, <a href=\"http://www.bourbaki.ens.fr/TEXTES/1108.pdf\" rel=\"nofollow noreferrer\">http://www.bourbaki.ens.fr/TEXTES/1108.pdf</a></p>\n</blockquote>\n" }, { "answer_id": 260475, "score": 7, "text": "<p>Thanks to @nfdc's comment following up on a comment of mine to the question <a href=\"https://mathoverflow.net/questions/260407/compactness-of-the-automorphic-quotient/260426\">Compactness of the automorphic quotient</a>, I/we should know that Sem. Bourb. is archived in the numdam.org at</p>\n\n<p><a href=\"http://www.numdam.org/actas/SB/\" rel=\"nofollow noreferrer\">http://www.numdam.org/actas/SB/</a></p>\n\n<p>EDIT: As @მამუკაჯიბლაძე observes in a comment, this includes the seminar up through 2006.</p>\n" }, { "answer_id": 393457, "score": 5, "text": "<p>Here you have a complete and updated set of resources\n<a href=\"https://www.bourbaki.fr/index-en.html\" rel=\"noreferrer\">https://www.bourbaki.fr/index-en.html</a></p>\n<ul>\n<li> Elements of mathematics <br>\nand, in particular, the two new volumes (Topologie Algébrique and Théories Spectrales) \n<li> Séminaire Betty B. <br>\nThe junior Bourbaki Seminar\n<li> N. Bourbaki Seminar <br>\nThe very one <br>\nPrevious sessions \n<li> Recorded talks of the séminaire N. Bourbaki\n(on youtube)\n<li> Earlier Texts <br>\nVolumes of the Seminar since its creation. <br>\nTexts of the Bourbaki Seminar lectures until 2010 (on Numdam) \n<li> Poincaré Seminar <br>\nIf you like Physics.\n</ul>\n" }, { "answer_id": 326938, "score": 1, "text": "<p>If you are interested in the actual talks, rather than the books, you can find it on youtube, <a href=\"https://www.youtube.com/user/PoincareInstitute/videos\" rel=\"nofollow noreferrer\">here</a>. They also stream them in direct, during the talks !</p>\n" }, { "answer_id": 326494, "score": 10, "text": "<p>This is an extension on Paul's answer, but it would have been too extensive of an edit so I posted this separately.</p>\n\n<p>(I have corrected the link in Paul's post since the old one was no longer working)</p>\n\n<blockquote>\n <p>Articles from <strong>1948</strong> (creation) up to <strong>2006</strong> can be found at <a href=\"http://www.numdam.org/actas/SB/\" rel=\"nofollow noreferrer\">this</a> link in Paul's post</p>\n \n <p>Articles <em>after</em> 2006 can also be purchased from SMF (Société Mathématique de France) at the following links: <a href=\"http://smf4.emath.fr/Publications/Asterisque/2008/317/html\" rel=\"nofollow noreferrer\">2006-2007</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2009/326/html\" rel=\"nofollow noreferrer\">2007-2008</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2010/332/html\" rel=\"nofollow noreferrer\">2008-2009</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2011/339/html\" rel=\"nofollow noreferrer\">2009-2010</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2012/348/html\" rel=\"nofollow noreferrer\">2010-2011</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2013/352/html\" rel=\"nofollow noreferrer\">2011-2012</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2014/361/html\" rel=\"nofollow noreferrer\">2012-2013</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2015/367-368/html\" rel=\"nofollow noreferrer\">2013-2014</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2016/380/html\" rel=\"nofollow noreferrer\">2014-2015</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2017/390/html\" rel=\"nofollow noreferrer\">2015-2016</a>, <a href=\"http://smf4.emath.fr/Publications/Asterisque/2017/390/html\" rel=\"nofollow noreferrer\">2016-2017</a>, (articles after 2017 can only be accessed, at the time of posting this, by clicking \"PDF\" next to the bolded article title at the following links: <a href=\"http://www.bourbaki.ens.fr/seminaires/2018/\" rel=\"nofollow noreferrer\">2018</a>, <a href=\"http://www.bourbaki.ens.fr/seminaires/2019/\" rel=\"nofollow noreferrer\">2019</a>).</p>\n</blockquote>\n\n<p><s>Note: Since the 2019 seminar is 3 days from the time of writing this, those files are not available at this time.</s></p>\n\n<p>As of April 1st, the articles from the March 2019 seminar are available at the link above.</p>\n\n<blockquote>\n <p>If you are wanting to download an article published <strong>after 2006</strong>, you just need to know the corresponding Exposé number which can be found under \"CONTENU DU VOLUME\" on the SMF links above, after which the article can be found at links with this format:\n <code>http://www.bourbaki.ens.fr/TEXTES/(put the exposé # here).pdf</code>. <br/>For example, <a href=\"http://www.bourbaki.ens.fr/TEXTES/1108.pdf\" rel=\"nofollow noreferrer\">http://www.bourbaki.ens.fr/TEXTES/1108.pdf</a></p>\n</blockquote>\n" }, { "answer_id": 260475, "score": 7, "text": "<p>Thanks to @nfdc's comment following up on a comment of mine to the question <a href=\"https://mathoverflow.net/questions/260407/compactness-of-the-automorphic-quotient/260426\">Compactness of the automorphic quotient</a>, I/we should know that Sem. Bourb. is archived in the numdam.org at</p>\n\n<p><a href=\"http://www.numdam.org/actas/SB/\" rel=\"nofollow noreferrer\">http://www.numdam.org/actas/SB/</a></p>\n\n<p>EDIT: As @მამუკაჯიბლაძე observes in a comment, this includes the seminar up through 2006.</p>\n" }, { "answer_id": 393457, "score": 5, "text": "<p>Here you have a complete and updated set of resources\n<a href=\"https://www.bourbaki.fr/index-en.html\" rel=\"noreferrer\">https://www.bourbaki.fr/index-en.html</a></p>\n<ul>\n<li> Elements of mathematics <br>\nand, in particular, the two new volumes (Topologie Algébrique and Théories Spectrales) \n<li> Séminaire Betty B. <br>\nThe junior Bourbaki Seminar\n<li> N. Bourbaki Seminar <br>\nThe very one <br>\nPrevious sessions \n<li> Recorded talks of the séminaire N. Bourbaki\n(on youtube)\n<li> Earlier Texts <br>\nVolumes of the Seminar since its creation. <br>\nTexts of the Bourbaki Seminar lectures until 2010 (on Numdam) \n<li> Poincaré Seminar <br>\nIf you like Physics.\n</ul>\n" }, { "answer_id": 326938, "score": 1, "text": "<p>If you are interested in the actual talks, rather than the books, you can find it on youtube, <a href=\"https://www.youtube.com/user/PoincareInstitute/videos\" rel=\"nofollow noreferrer\">here</a>. They also stream them in direct, during the talks !</p>\n" } ]
Question: Where is Seminaire Bourbaki on-line? Answer 1: <p>This is an extension on Paul's answer, but it would have been too extensive of an edit so I posted this separately.</p> <p>(I have corrected the link in Paul's post since the old one was no longer working)</p> <blockquote> <p>Articles from <strong>1948</strong> (creation) up to <strong>2006</strong> can be found at <a href="http://www.numdam.org/actas/SB/" rel="nofollow noreferrer">this</a> link in Paul's post</p> <p>Articles <em>after</em> 2006 can also be purchased from SMF (Société Mathématique de France) at the following links: <a href="http://smf4.emath.fr/Publications/Asterisque/2008/317/html" rel="nofollow noreferrer">2006-2007</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2009/326/html" rel="nofollow noreferrer">2007-2008</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2010/332/html" rel="nofollow noreferrer">2008-2009</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2011/339/html" rel="nofollow noreferrer">2009-2010</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2012/348/html" rel="nofollow noreferrer">2010-2011</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2013/352/html" rel="nofollow noreferrer">2011-2012</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2014/361/html" rel="nofollow noreferrer">2012-2013</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2015/367-368/html" rel="nofollow noreferrer">2013-2014</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2016/380/html" rel="nofollow noreferrer">2014-2015</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2017/390/html" rel="nofollow noreferrer">2015-2016</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2017/390/html" rel="nofollow noreferrer">2016-2017</a>, (articles after 2017 can only be accessed, at the time of posting this, by clicking "PDF" next to the bolded article title at the following links: <a href="http://www.bourbaki.ens.fr/seminaires/2018/" rel="nofollow noreferrer">2018</a>, <a href="http://www.bourbaki.ens.fr/seminaires/2019/" rel="nofollow noreferrer">2019</a>).</p> </blockquote> <p><s>Note: Since the 2019 seminar is 3 days from the time of writing this, those files are not available at this time.</s></p> <p>As of April 1st, the articles from the March 2019 seminar are available at the link above.</p> <blockquote> <p>If you are wanting to download an article published <strong>after 2006</strong>, you just need to know the corresponding Exposé number which can be found under "CONTENU DU VOLUME" on the SMF links above, after which the article can be found at links with this format: <code>http://www.bourbaki.ens.fr/TEXTES/(put the exposé # here).pdf</code>. <br/>For example, <a href="http://www.bourbaki.ens.fr/TEXTES/1108.pdf" rel="nofollow noreferrer">http://www.bourbaki.ens.fr/TEXTES/1108.pdf</a></p> </blockquote> Answer 2: <p>Thanks to @nfdc's comment following up on a comment of mine to the question <a href="https://mathoverflow.net/questions/260407/compactness-of-the-automorphic-quotient/260426">Compactness of the automorphic quotient</a>, I/we should know that Sem. Bourb. is archived in the numdam.org at</p> <p><a href="http://www.numdam.org/actas/SB/" rel="nofollow noreferrer">http://www.numdam.org/actas/SB/</a></p> <p>EDIT: As @მამუკაჯიბლაძე observes in a comment, this includes the seminar up through 2006.</p> Answer 3: <p>Here you have a complete and updated set of resources <a href="https://www.bourbaki.fr/index-en.html" rel="noreferrer">https://www.bourbaki.fr/index-en.html</a></p> <ul> <li> Elements of mathematics <br> and, in particular, the two new volumes (Topologie Algébrique and Théories Spectrales) <li> Séminaire Betty B. <br> The junior Bourbaki Seminar <li> N. Bourbaki Seminar <br> The very one <br> Previous sessions <li> Recorded talks of the séminaire N. Bourbaki (on youtube) <li> Earlier Texts <br> Volumes of the Seminar since its creation. <br> Texts of the Bourbaki Seminar lectures until 2010 (on Numdam) <li> Poincaré Seminar <br> If you like Physics. </ul> Answer 4: <p>If you are interested in the actual talks, rather than the books, you can find it on youtube, <a href="https://www.youtube.com/user/PoincareInstitute/videos" rel="nofollow noreferrer">here</a>. They also stream them in direct, during the talks !</p> Answer 5: <p>This is an extension on Paul's answer, but it would have been too extensive of an edit so I posted this separately.</p> <p>(I have corrected the link in Paul's post since the old one was no longer working)</p> <blockquote> <p>Articles from <strong>1948</strong> (creation) up to <strong>2006</strong> can be found at <a href="http://www.numdam.org/actas/SB/" rel="nofollow noreferrer">this</a> link in Paul's post</p> <p>Articles <em>after</em> 2006 can also be purchased from SMF (Société Mathématique de France) at the following links: <a href="http://smf4.emath.fr/Publications/Asterisque/2008/317/html" rel="nofollow noreferrer">2006-2007</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2009/326/html" rel="nofollow noreferrer">2007-2008</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2010/332/html" rel="nofollow noreferrer">2008-2009</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2011/339/html" rel="nofollow noreferrer">2009-2010</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2012/348/html" rel="nofollow noreferrer">2010-2011</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2013/352/html" rel="nofollow noreferrer">2011-2012</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2014/361/html" rel="nofollow noreferrer">2012-2013</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2015/367-368/html" rel="nofollow noreferrer">2013-2014</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2016/380/html" rel="nofollow noreferrer">2014-2015</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2017/390/html" rel="nofollow noreferrer">2015-2016</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2017/390/html" rel="nofollow noreferrer">2016-2017</a>, (articles after 2017 can only be accessed, at the time of posting this, by clicking "PDF" next to the bolded article title at the following links: <a href="http://www.bourbaki.ens.fr/seminaires/2018/" rel="nofollow noreferrer">2018</a>, <a href="http://www.bourbaki.ens.fr/seminaires/2019/" rel="nofollow noreferrer">2019</a>).</p> </blockquote> <p><s>Note: Since the 2019 seminar is 3 days from the time of writing this, those files are not available at this time.</s></p> <p>As of April 1st, the articles from the March 2019 seminar are available at the link above.</p> <blockquote> <p>If you are wanting to download an article published <strong>after 2006</strong>, you just need to know the corresponding Exposé number which can be found under "CONTENU DU VOLUME" on the SMF links above, after which the article can be found at links with this format: <code>http://www.bourbaki.ens.fr/TEXTES/(put the exposé # here).pdf</code>. <br/>For example, <a href="http://www.bourbaki.ens.fr/TEXTES/1108.pdf" rel="nofollow noreferrer">http://www.bourbaki.ens.fr/TEXTES/1108.pdf</a></p> </blockquote> Answer 6: <p>Thanks to @nfdc's comment following up on a comment of mine to the question <a href="https://mathoverflow.net/questions/260407/compactness-of-the-automorphic-quotient/260426">Compactness of the automorphic quotient</a>, I/we should know that Sem. Bourb. is archived in the numdam.org at</p> <p><a href="http://www.numdam.org/actas/SB/" rel="nofollow noreferrer">http://www.numdam.org/actas/SB/</a></p> <p>EDIT: As @მამუკაჯიბლაძე observes in a comment, this includes the seminar up through 2006.</p> Answer 7: <p>Here you have a complete and updated set of resources <a href="https://www.bourbaki.fr/index-en.html" rel="noreferrer">https://www.bourbaki.fr/index-en.html</a></p> <ul> <li> Elements of mathematics <br> and, in particular, the two new volumes (Topologie Algébrique and Théories Spectrales) <li> Séminaire Betty B. <br> The junior Bourbaki Seminar <li> N. Bourbaki Seminar <br> The very one <br> Previous sessions <li> Recorded talks of the séminaire N. Bourbaki (on youtube) <li> Earlier Texts <br> Volumes of the Seminar since its creation. <br> Texts of the Bourbaki Seminar lectures until 2010 (on Numdam) <li> Poincaré Seminar <br> If you like Physics. </ul> Answer 8: <p>If you are interested in the actual talks, rather than the books, you can find it on youtube, <a href="https://www.youtube.com/user/PoincareInstitute/videos" rel="nofollow noreferrer">here</a>. They also stream them in direct, during the talks !</p> Answer 9: <p>This is an extension on Paul's answer, but it would have been too extensive of an edit so I posted this separately.</p> <p>(I have corrected the link in Paul's post since the old one was no longer working)</p> <blockquote> <p>Articles from <strong>1948</strong> (creation) up to <strong>2006</strong> can be found at <a href="http://www.numdam.org/actas/SB/" rel="nofollow noreferrer">this</a> link in Paul's post</p> <p>Articles <em>after</em> 2006 can also be purchased from SMF (Société Mathématique de France) at the following links: <a href="http://smf4.emath.fr/Publications/Asterisque/2008/317/html" rel="nofollow noreferrer">2006-2007</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2009/326/html" rel="nofollow noreferrer">2007-2008</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2010/332/html" rel="nofollow noreferrer">2008-2009</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2011/339/html" rel="nofollow noreferrer">2009-2010</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2012/348/html" rel="nofollow noreferrer">2010-2011</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2013/352/html" rel="nofollow noreferrer">2011-2012</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2014/361/html" rel="nofollow noreferrer">2012-2013</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2015/367-368/html" rel="nofollow noreferrer">2013-2014</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2016/380/html" rel="nofollow noreferrer">2014-2015</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2017/390/html" rel="nofollow noreferrer">2015-2016</a>, <a href="http://smf4.emath.fr/Publications/Asterisque/2017/390/html" rel="nofollow noreferrer">2016-2017</a>, (articles after 2017 can only be accessed, at the time of posting this, by clicking "PDF" next to the bolded article title at the following links: <a href="http://www.bourbaki.ens.fr/seminaires/2018/" rel="nofollow noreferrer">2018</a>, <a href="http://www.bourbaki.ens.fr/seminaires/2019/" rel="nofollow noreferrer">2019</a>).</p> </blockquote> <p><s>Note: Since the 2019 seminar is 3 days from the time of writing this, those files are not available at this time.</s></p> <p>As of April 1st, the articles from the March 2019 seminar are available at the link above.</p> <blockquote> <p>If you are wanting to download an article published <strong>after 2006</strong>, you just need to know the corresponding Exposé number which can be found under "CONTENU DU VOLUME" on the SMF links above, after which the article can be found at links with this format: <code>http://www.bourbaki.ens.fr/TEXTES/(put the exposé # here).pdf</code>. <br/>For example, <a href="http://www.bourbaki.ens.fr/TEXTES/1108.pdf" rel="nofollow noreferrer">http://www.bourbaki.ens.fr/TEXTES/1108.pdf</a></p> </blockquote> Answer 10: <p>Thanks to @nfdc's comment following up on a comment of mine to the question <a href="https://mathoverflow.net/questions/260407/compactness-of-the-automorphic-quotient/260426">Compactness of the automorphic quotient</a>, I/we should know that Sem. Bourb. is archived in the numdam.org at</p> <p><a href="http://www.numdam.org/actas/SB/" rel="nofollow noreferrer">http://www.numdam.org/actas/SB/</a></p> <p>EDIT: As @მამუკაჯიბლაძე observes in a comment, this includes the seminar up through 2006.</p> Answer 11: <p>Here you have a complete and updated set of resources <a href="https://www.bourbaki.fr/index-en.html" rel="noreferrer">https://www.bourbaki.fr/index-en.html</a></p> <ul> <li> Elements of mathematics <br> and, in particular, the two new volumes (Topologie Algébrique and Théories Spectrales) <li> Séminaire Betty B. <br> The junior Bourbaki Seminar <li> N. Bourbaki Seminar <br> The very one <br> Previous sessions <li> Recorded talks of the séminaire N. Bourbaki (on youtube) <li> Earlier Texts <br> Volumes of the Seminar since its creation. <br> Texts of the Bourbaki Seminar lectures until 2010 (on Numdam) <li> Poincaré Seminar <br> If you like Physics. </ul> Answer 12: <p>If you are interested in the actual talks, rather than the books, you can find it on youtube, <a href="https://www.youtube.com/user/PoincareInstitute/videos" rel="nofollow noreferrer">here</a>. They also stream them in direct, during the talks !</p>
Weber&#39;s class number problem and real quadratic fields of class number one
258,672
10
https://mathoverflow.net/questions/258672/webers-class-number-problem-and-real-quadratic-fields-of-class-number-one
<p>(A) It is an old, outstanding problem to show that there are infinitely many real quadratic fields with class number one.</p> <p>(B) On the other hand, Weber's class number problem (for $p=2$) asks to show that the degree $2^n$ cyclotomic extension $F_n=\mathbb Q(\cos(2\pi/2^{n+2}))$ of $\mathbb Q$ has class number one, for all $n$.</p> <p>According to p.41 of <a href="http://wwwf.imperial.ac.uk/~buzzard/maths/research/notes/Yukako_Kezuka_MSc_Project.pdf" rel="noreferrer">this thesis</a>, showing (B) for infinitely many $n$ implies (A), but no idea of the proof is given. This does not seem to follow immediately for me, but perhaps I am missing something simple here. Certainly $F_n$ contains a real quadratic subfield, but why should its class number also be 1?</p>
[ { "answer_id": 258757, "score": 9, "text": "<p>This statement is most certainly nonsense; I guess what she meant to write was that in order to prove the existence of infinitely many number fields with class number $1$ it is sufficient to prove $h(F_n) = 1$ for infinitely many (and therefore for all) $n$. </p>\n" }, { "answer_id": 258673, "score": 4, "text": "<p>Let $K$ be any subextension of $F_n$. Then the natural map of class groups $\\text{Cl}(K) \\rightarrow \\text{Cl}(F_n)$ is an injection. Indeed, you can see it in terms of unramified abelian extension via global class field theory. If $L$ is an unramified abelian extension of $K$, since $F_n$ is totally ramified at primes over $2$ in $K$, we have $L \\cap F_n = K$. Therefore the compositum $L \\cdot F_n$ has Galois group $\\text{Gal}(L/K)$ over $F_n$.</p>\n" }, { "answer_id": 258757, "score": 9, "text": "<p>This statement is most certainly nonsense; I guess what she meant to write was that in order to prove the existence of infinitely many number fields with class number $1$ it is sufficient to prove $h(F_n) = 1$ for infinitely many (and therefore for all) $n$. </p>\n" }, { "answer_id": 258673, "score": 4, "text": "<p>Let $K$ be any subextension of $F_n$. Then the natural map of class groups $\\text{Cl}(K) \\rightarrow \\text{Cl}(F_n)$ is an injection. Indeed, you can see it in terms of unramified abelian extension via global class field theory. If $L$ is an unramified abelian extension of $K$, since $F_n$ is totally ramified at primes over $2$ in $K$, we have $L \\cap F_n = K$. Therefore the compositum $L \\cdot F_n$ has Galois group $\\text{Gal}(L/K)$ over $F_n$.</p>\n" }, { "answer_id": 258757, "score": 9, "text": "<p>This statement is most certainly nonsense; I guess what she meant to write was that in order to prove the existence of infinitely many number fields with class number $1$ it is sufficient to prove $h(F_n) = 1$ for infinitely many (and therefore for all) $n$. </p>\n" }, { "answer_id": 258673, "score": 4, "text": "<p>Let $K$ be any subextension of $F_n$. Then the natural map of class groups $\\text{Cl}(K) \\rightarrow \\text{Cl}(F_n)$ is an injection. Indeed, you can see it in terms of unramified abelian extension via global class field theory. If $L$ is an unramified abelian extension of $K$, since $F_n$ is totally ramified at primes over $2$ in $K$, we have $L \\cap F_n = K$. Therefore the compositum $L \\cdot F_n$ has Galois group $\\text{Gal}(L/K)$ over $F_n$.</p>\n" } ]
Question: Weber&#39;s class number problem and real quadratic fields of class number one Answer 1: <p>This statement is most certainly nonsense; I guess what she meant to write was that in order to prove the existence of infinitely many number fields with class number $1$ it is sufficient to prove $h(F_n) = 1$ for infinitely many (and therefore for all) $n$. </p> Answer 2: <p>Let $K$ be any subextension of $F_n$. Then the natural map of class groups $\text{Cl}(K) \rightarrow \text{Cl}(F_n)$ is an injection. Indeed, you can see it in terms of unramified abelian extension via global class field theory. If $L$ is an unramified abelian extension of $K$, since $F_n$ is totally ramified at primes over $2$ in $K$, we have $L \cap F_n = K$. Therefore the compositum $L \cdot F_n$ has Galois group $\text{Gal}(L/K)$ over $F_n$.</p> Answer 3: <p>This statement is most certainly nonsense; I guess what she meant to write was that in order to prove the existence of infinitely many number fields with class number $1$ it is sufficient to prove $h(F_n) = 1$ for infinitely many (and therefore for all) $n$. </p> Answer 4: <p>Let $K$ be any subextension of $F_n$. Then the natural map of class groups $\text{Cl}(K) \rightarrow \text{Cl}(F_n)$ is an injection. Indeed, you can see it in terms of unramified abelian extension via global class field theory. If $L$ is an unramified abelian extension of $K$, since $F_n$ is totally ramified at primes over $2$ in $K$, we have $L \cap F_n = K$. Therefore the compositum $L \cdot F_n$ has Galois group $\text{Gal}(L/K)$ over $F_n$.</p> Answer 5: <p>This statement is most certainly nonsense; I guess what she meant to write was that in order to prove the existence of infinitely many number fields with class number $1$ it is sufficient to prove $h(F_n) = 1$ for infinitely many (and therefore for all) $n$. </p> Answer 6: <p>Let $K$ be any subextension of $F_n$. Then the natural map of class groups $\text{Cl}(K) \rightarrow \text{Cl}(F_n)$ is an injection. Indeed, you can see it in terms of unramified abelian extension via global class field theory. If $L$ is an unramified abelian extension of $K$, since $F_n$ is totally ramified at primes over $2$ in $K$, we have $L \cap F_n = K$. Therefore the compositum $L \cdot F_n$ has Galois group $\text{Gal}(L/K)$ over $F_n$.</p>
Does the Hasse norm theorem easily imply the global squares theorem?
259,395
10
https://mathoverflow.net/questions/259395/does-the-hasse-norm-theorem-easily-imply-the-global-squares-theorem
<p>The Hasse norm theorem says a nonzero element $x$ of a number field $K$ is a norm for a cyclic extension $L/K$ iff $x$ is a norm for all completions $L_p/K_p$ at primes $p$ of $L$ (writing $p$ also for the restriction to $K$).</p> <p>The global squares theorem says $x$ is a square in number field $K$ iff it is a square in every completion $K_p$ at a prime of $K$. </p> <p>Is there some quick proof of the global squares theorem from the global norm theorem? Hilbert proved the global norm theorem for the case where $L/K$ is a quadratic extension. Is that special case enough to quickly prove the global squares theorem?</p> <p>I would especially like to know a constructive proof that if $x$ is not a square in number field $K$ then there is a quadratic extension $L/K$ for which $x$ is not a norm. </p>
[ { "answer_id": 259639, "score": 4, "text": "<p>I'm not sure how constructive the following answer to your second question is.\nIf $\\alpha$ is not a square, then there must exist a prime ideal ${\\mathfrak p}$\ncoprime to $2\\alpha$ with $(\\alpha/{\\mathfrak p}) = -1$. This follows from Kronecker's density theorem according to which the density of the prime ideals that split completely in a normal extension $L/K$ is equal to $1/(L:K)$. All you need for a proof is that the Dedekind zeta function of a number field $F$ has the property that $(s-1) \\zeta_F(s)$ has a finite limit as $s \\to 1$. </p>\n\n<p>Once you have such a prime ideal ${\\mathfrak p}$, take any quadratic extension $L/K$ in which ${\\mathfrak p}$ ramifies (you might pick an ideal ${\\mathfrak a}$\nin the inverse class, write ${\\mathfrak p}{\\mathfrak a} = (\\beta)$, and set $L = K(\\sqrt{\\beta})$, for examples), and $\\alpha$ will not be a norm from $L$.</p>\n\n<p>I would be suprised if you could construct ${\\mathfrak p}$ \"explicitly\". Take e.g. the case where $\\sqrt{\\alpha}$ generates the Hilbert class field of a number field $K$ with class number $2$; then you would have to find a nonprincipal prime ideal in $K$.The fastest solution in this case is, as far as I know, picking prime ideals at random.</p>\n" }, { "answer_id": 259648, "score": 2, "text": "<p>Well, I would say that this crucially depends on what you define to be \"quick\". If you admit global class field theory, at least in its idelic formulation, the fact that all primes in $K$ are split in the extensions $K(\\sqrt{x})/K$ proves that $x$ is a square quite rapidly (as Timo Keller has observed in the comments, this is actually easier than Chebotarev; and Franz Lemmermeyer's answer says even more). I give a short proof at the end of my answer, although it does not \"need\" the global norm theorem (which is anyhow hidden beneath the idelic formulation of global class field theory). I am pretty sure that this is the proof you have in mind when you ask for a <em>quicker one</em>.</p>\n\n<p>Without class field theory, I could come up only with an \"elementary\" proof assuming that </p>\n\n<p><strong>[H1]</strong> $K$ admits at least one real embedding, let it be $\\sigma\\colon K\\hookrightarrow \\mathbb{R}$, and</p>\n\n<p><strong>[H2]</strong> $K/\\mathbb{Q}$ is unramified at $2$.</p>\n\n<p>Consider then the following</p>\n\n<p><strong>Claim</strong> <em>An element in $\\mathcal{O}_K\\setminus\\{0\\}$ is a square if and only if it is a (global) norm from every quadratic extension $L/K$</em></p>\n\n<p>Admit the claim: we are going to deduce that if an element $x=r/s$ with $r,s\\in\\mathcal{O}_K$ is locally a square everywhere, then $x\\in(K^\\times)^2$. Let $y=xs^2$: then $y\\in\\mathcal{O}_K\\setminus\\{0\\}$ is locally a norm from every quadratic extension $L/K$, since it is a square in $K_\\mathfrak{p}^\\times$ for all $\\mathfrak{p}$. By the global norm theorem it is a global norm from every quadratic extension, and the claim tells you $y\\in (K^\\times)^2$: hence the same holds for $x=y/s^2$.</p>\n\n<p>To prove the claim, observe that one direction is obvious: squares are norms from every quadratic extension. On the other hand, let $x\\in\\mathcal{O}_K\\setminus\\{0\\}$ be a norm from every quadratic extension. I will use the real embedding $\\sigma$. Consider the function\n$$\nf\\colon \\mathcal{O}_K\\times\\mathcal{O}_K\\setminus \\{0\\}\\longrightarrow \\mathbb{R}\n$$\nsending $(a,b)$ to $f(a,b)=\\sigma\\left(\\left(\\frac{a}{b}\\right)^2-\\frac{x}{b^2}\\right)$. Then there is a bound $B\\in\\mathbb{R}$ such that $\\mathrm{Im}(f)\\subseteq [B,+\\infty)$. Indeed, $(a/b)^2\\geq 0$, while $b\\mapsto x/b^2$ is bounded from below, since $\\sigma(\\mathcal{O}_K)$ is discrete. Chose now a prime number $\\ell\\equiv 3\\pmod{4}$ which is unramfied in $K/\\mathbb{Q}$ and such that $-\\ell&lt;B$. Consider the extension $K(\\sqrt{-\\ell})/K$: since the discriminant of $\\mathcal{O}_K(\\sqrt{-\\ell})/\\mathcal{O}_K$ is $-4\\ell$ and the ideal $\\ell\\mathcal{O}_K$ is square-free by assumption, it follows that $[\\mathcal{O}_L:\\mathcal{O}_K(\\sqrt{-\\ell})]=1,2$. Since we assumed that $K/\\mathbb{Q}$ be unramified at $2$, the same argument as above implies that $\\mathcal{O}_{K(i)}=\\mathcal{O}_K[i]$ and then that $x^2+y^2\\equiv 0\\pmod{4\\mathcal{O}_K}$ has the unique solution $x\\equiv y\\equiv 0\\pmod{4\\mathcal{O}_K}$ (because each prime $\\mathfrak{r}\\mid 2$ of $\\mathcal{O}_K$ is ramified in $K(i)/K$, hence if something has norm from $K(i)$ to $K$ which is divisible by $4$, it must lie in $2\\mathcal{O}_K[i]$): in particular, the usual proof that the ring of integers of the imaginary quadratic field $\\mathbb{Q}(\\sqrt{-\\ell})$ is $\\mathbb{Z}[\\sqrt{-\\ell}]$ carries over, and $[\\mathcal{O}_L:\\mathcal{O}_K(\\sqrt{-\\ell})]=1$.</p>\n\n<p>Since we assumed that $x$ is a norm from $K(\\sqrt{-\\ell})/K$, we get $x=a^2+\\ell b^2$ for suitable $a,b\\in\\mathcal{O}_K$, by the above discussion. If $b=0$, then $x$ is a square, and we win. If not, rewrite the equality as\n$$\n\\frac{a^2-x}{b^2}=-\\ell\\Longrightarrow \\sigma(-\\ell)=-\\ell=f(a,b)\n$$\nThis contradicts the fact that $\\mathrm(Im)(f)\\subseteq [B,+\\infty)$ and finishes the proof of the claim, but I realise that one can reasonably object that the above proof is not \"quick\"...</p>\n\n<p>One final comment about assumption <strong>[H2]</strong>. It was only needed above to be sure that $\\mathcal{O}_{L}=\\mathcal{O}_K[\\sqrt{-\\ell}]$. This is not necessary: one could do, for instance, by proving that there must exist a $\\delta$ with $\\sigma(-\\delta)&lt;B$ and $\\mathcal{O}_{K(\\sqrt{-\\delta})}=\\mathcal{O}_K[\\sqrt{-\\delta}]$. Or, if one assumes that $2\\notin (K^\\times)^2$, then also the case $[\\mathcal{O}_{L}\\colon\\mathcal{O}_K]=2$ can be treated, by modifying a bit $f(a,b)$. But I could not do much better, hence I left assumption <strong>[H2]</strong>, although probably too restrective.</p>\n\n<p>Now, with global class field theory. The extension $L=K(\\sqrt{x})/K$ is abelian and corresponds to a unique normic subgroup $\\mathcal{H}\\subseteq C_K$ where $C_K$ is the idèle class group of $K$, so $\\mathcal{H}=\\mathrm{Norm}_{L/K}(C_L)$. Your assumption that $x\\in (K_\\mathfrak{p}^\\times)^2$ for all primes $\\mathfrak{p}$ means that $\\mathrm{Norm}_{L_\\mathfrak{P}/K_\\mathfrak{p}}(L_\\mathfrak{P}^\\times)=K_\\mathfrak{p}^\\times$ for all $\\mathfrak{p}$ and all $\\mathfrak{P}\\mid\\mathfrak{p}$ (<em>caution: we are using here that the minimal polynomial for $x$ has degree $1$ or $2$. Indeed, saying that $x\\in (K_\\mathfrak{p}^\\times)^2$ is equivalent to saying that all roots of $X^2-x$ are in $K_\\mathfrak{p}$, but this can be false if we replace $X^2-x$ by $X^n-x$ and try to prove the analogous statement replacing \"square\" by \"$n$th power\"</em>). Then $\\mathrm{Norm}_{L/K}(C_L)=C_K$ and hence $L=K$, or $x\\in(K^\\times)^2$. This is an adaptation of Corollary 8.8 in Tate's article on Global Class Field Theory in Cassels and Fröhlich's <em>Algebraic Number Theory</em>.</p>\n" }, { "answer_id": 259639, "score": 4, "text": "<p>I'm not sure how constructive the following answer to your second question is.\nIf $\\alpha$ is not a square, then there must exist a prime ideal ${\\mathfrak p}$\ncoprime to $2\\alpha$ with $(\\alpha/{\\mathfrak p}) = -1$. This follows from Kronecker's density theorem according to which the density of the prime ideals that split completely in a normal extension $L/K$ is equal to $1/(L:K)$. All you need for a proof is that the Dedekind zeta function of a number field $F$ has the property that $(s-1) \\zeta_F(s)$ has a finite limit as $s \\to 1$. </p>\n\n<p>Once you have such a prime ideal ${\\mathfrak p}$, take any quadratic extension $L/K$ in which ${\\mathfrak p}$ ramifies (you might pick an ideal ${\\mathfrak a}$\nin the inverse class, write ${\\mathfrak p}{\\mathfrak a} = (\\beta)$, and set $L = K(\\sqrt{\\beta})$, for examples), and $\\alpha$ will not be a norm from $L$.</p>\n\n<p>I would be suprised if you could construct ${\\mathfrak p}$ \"explicitly\". Take e.g. the case where $\\sqrt{\\alpha}$ generates the Hilbert class field of a number field $K$ with class number $2$; then you would have to find a nonprincipal prime ideal in $K$.The fastest solution in this case is, as far as I know, picking prime ideals at random.</p>\n" }, { "answer_id": 259648, "score": 2, "text": "<p>Well, I would say that this crucially depends on what you define to be \"quick\". If you admit global class field theory, at least in its idelic formulation, the fact that all primes in $K$ are split in the extensions $K(\\sqrt{x})/K$ proves that $x$ is a square quite rapidly (as Timo Keller has observed in the comments, this is actually easier than Chebotarev; and Franz Lemmermeyer's answer says even more). I give a short proof at the end of my answer, although it does not \"need\" the global norm theorem (which is anyhow hidden beneath the idelic formulation of global class field theory). I am pretty sure that this is the proof you have in mind when you ask for a <em>quicker one</em>.</p>\n\n<p>Without class field theory, I could come up only with an \"elementary\" proof assuming that </p>\n\n<p><strong>[H1]</strong> $K$ admits at least one real embedding, let it be $\\sigma\\colon K\\hookrightarrow \\mathbb{R}$, and</p>\n\n<p><strong>[H2]</strong> $K/\\mathbb{Q}$ is unramified at $2$.</p>\n\n<p>Consider then the following</p>\n\n<p><strong>Claim</strong> <em>An element in $\\mathcal{O}_K\\setminus\\{0\\}$ is a square if and only if it is a (global) norm from every quadratic extension $L/K$</em></p>\n\n<p>Admit the claim: we are going to deduce that if an element $x=r/s$ with $r,s\\in\\mathcal{O}_K$ is locally a square everywhere, then $x\\in(K^\\times)^2$. Let $y=xs^2$: then $y\\in\\mathcal{O}_K\\setminus\\{0\\}$ is locally a norm from every quadratic extension $L/K$, since it is a square in $K_\\mathfrak{p}^\\times$ for all $\\mathfrak{p}$. By the global norm theorem it is a global norm from every quadratic extension, and the claim tells you $y\\in (K^\\times)^2$: hence the same holds for $x=y/s^2$.</p>\n\n<p>To prove the claim, observe that one direction is obvious: squares are norms from every quadratic extension. On the other hand, let $x\\in\\mathcal{O}_K\\setminus\\{0\\}$ be a norm from every quadratic extension. I will use the real embedding $\\sigma$. Consider the function\n$$\nf\\colon \\mathcal{O}_K\\times\\mathcal{O}_K\\setminus \\{0\\}\\longrightarrow \\mathbb{R}\n$$\nsending $(a,b)$ to $f(a,b)=\\sigma\\left(\\left(\\frac{a}{b}\\right)^2-\\frac{x}{b^2}\\right)$. Then there is a bound $B\\in\\mathbb{R}$ such that $\\mathrm{Im}(f)\\subseteq [B,+\\infty)$. Indeed, $(a/b)^2\\geq 0$, while $b\\mapsto x/b^2$ is bounded from below, since $\\sigma(\\mathcal{O}_K)$ is discrete. Chose now a prime number $\\ell\\equiv 3\\pmod{4}$ which is unramfied in $K/\\mathbb{Q}$ and such that $-\\ell&lt;B$. Consider the extension $K(\\sqrt{-\\ell})/K$: since the discriminant of $\\mathcal{O}_K(\\sqrt{-\\ell})/\\mathcal{O}_K$ is $-4\\ell$ and the ideal $\\ell\\mathcal{O}_K$ is square-free by assumption, it follows that $[\\mathcal{O}_L:\\mathcal{O}_K(\\sqrt{-\\ell})]=1,2$. Since we assumed that $K/\\mathbb{Q}$ be unramified at $2$, the same argument as above implies that $\\mathcal{O}_{K(i)}=\\mathcal{O}_K[i]$ and then that $x^2+y^2\\equiv 0\\pmod{4\\mathcal{O}_K}$ has the unique solution $x\\equiv y\\equiv 0\\pmod{4\\mathcal{O}_K}$ (because each prime $\\mathfrak{r}\\mid 2$ of $\\mathcal{O}_K$ is ramified in $K(i)/K$, hence if something has norm from $K(i)$ to $K$ which is divisible by $4$, it must lie in $2\\mathcal{O}_K[i]$): in particular, the usual proof that the ring of integers of the imaginary quadratic field $\\mathbb{Q}(\\sqrt{-\\ell})$ is $\\mathbb{Z}[\\sqrt{-\\ell}]$ carries over, and $[\\mathcal{O}_L:\\mathcal{O}_K(\\sqrt{-\\ell})]=1$.</p>\n\n<p>Since we assumed that $x$ is a norm from $K(\\sqrt{-\\ell})/K$, we get $x=a^2+\\ell b^2$ for suitable $a,b\\in\\mathcal{O}_K$, by the above discussion. If $b=0$, then $x$ is a square, and we win. If not, rewrite the equality as\n$$\n\\frac{a^2-x}{b^2}=-\\ell\\Longrightarrow \\sigma(-\\ell)=-\\ell=f(a,b)\n$$\nThis contradicts the fact that $\\mathrm(Im)(f)\\subseteq [B,+\\infty)$ and finishes the proof of the claim, but I realise that one can reasonably object that the above proof is not \"quick\"...</p>\n\n<p>One final comment about assumption <strong>[H2]</strong>. It was only needed above to be sure that $\\mathcal{O}_{L}=\\mathcal{O}_K[\\sqrt{-\\ell}]$. This is not necessary: one could do, for instance, by proving that there must exist a $\\delta$ with $\\sigma(-\\delta)&lt;B$ and $\\mathcal{O}_{K(\\sqrt{-\\delta})}=\\mathcal{O}_K[\\sqrt{-\\delta}]$. Or, if one assumes that $2\\notin (K^\\times)^2$, then also the case $[\\mathcal{O}_{L}\\colon\\mathcal{O}_K]=2$ can be treated, by modifying a bit $f(a,b)$. But I could not do much better, hence I left assumption <strong>[H2]</strong>, although probably too restrective.</p>\n\n<p>Now, with global class field theory. The extension $L=K(\\sqrt{x})/K$ is abelian and corresponds to a unique normic subgroup $\\mathcal{H}\\subseteq C_K$ where $C_K$ is the idèle class group of $K$, so $\\mathcal{H}=\\mathrm{Norm}_{L/K}(C_L)$. Your assumption that $x\\in (K_\\mathfrak{p}^\\times)^2$ for all primes $\\mathfrak{p}$ means that $\\mathrm{Norm}_{L_\\mathfrak{P}/K_\\mathfrak{p}}(L_\\mathfrak{P}^\\times)=K_\\mathfrak{p}^\\times$ for all $\\mathfrak{p}$ and all $\\mathfrak{P}\\mid\\mathfrak{p}$ (<em>caution: we are using here that the minimal polynomial for $x$ has degree $1$ or $2$. Indeed, saying that $x\\in (K_\\mathfrak{p}^\\times)^2$ is equivalent to saying that all roots of $X^2-x$ are in $K_\\mathfrak{p}$, but this can be false if we replace $X^2-x$ by $X^n-x$ and try to prove the analogous statement replacing \"square\" by \"$n$th power\"</em>). Then $\\mathrm{Norm}_{L/K}(C_L)=C_K$ and hence $L=K$, or $x\\in(K^\\times)^2$. This is an adaptation of Corollary 8.8 in Tate's article on Global Class Field Theory in Cassels and Fröhlich's <em>Algebraic Number Theory</em>.</p>\n" }, { "answer_id": 259639, "score": 4, "text": "<p>I'm not sure how constructive the following answer to your second question is.\nIf $\\alpha$ is not a square, then there must exist a prime ideal ${\\mathfrak p}$\ncoprime to $2\\alpha$ with $(\\alpha/{\\mathfrak p}) = -1$. This follows from Kronecker's density theorem according to which the density of the prime ideals that split completely in a normal extension $L/K$ is equal to $1/(L:K)$. All you need for a proof is that the Dedekind zeta function of a number field $F$ has the property that $(s-1) \\zeta_F(s)$ has a finite limit as $s \\to 1$. </p>\n\n<p>Once you have such a prime ideal ${\\mathfrak p}$, take any quadratic extension $L/K$ in which ${\\mathfrak p}$ ramifies (you might pick an ideal ${\\mathfrak a}$\nin the inverse class, write ${\\mathfrak p}{\\mathfrak a} = (\\beta)$, and set $L = K(\\sqrt{\\beta})$, for examples), and $\\alpha$ will not be a norm from $L$.</p>\n\n<p>I would be suprised if you could construct ${\\mathfrak p}$ \"explicitly\". Take e.g. the case where $\\sqrt{\\alpha}$ generates the Hilbert class field of a number field $K$ with class number $2$; then you would have to find a nonprincipal prime ideal in $K$.The fastest solution in this case is, as far as I know, picking prime ideals at random.</p>\n" }, { "answer_id": 259648, "score": 2, "text": "<p>Well, I would say that this crucially depends on what you define to be \"quick\". If you admit global class field theory, at least in its idelic formulation, the fact that all primes in $K$ are split in the extensions $K(\\sqrt{x})/K$ proves that $x$ is a square quite rapidly (as Timo Keller has observed in the comments, this is actually easier than Chebotarev; and Franz Lemmermeyer's answer says even more). I give a short proof at the end of my answer, although it does not \"need\" the global norm theorem (which is anyhow hidden beneath the idelic formulation of global class field theory). I am pretty sure that this is the proof you have in mind when you ask for a <em>quicker one</em>.</p>\n\n<p>Without class field theory, I could come up only with an \"elementary\" proof assuming that </p>\n\n<p><strong>[H1]</strong> $K$ admits at least one real embedding, let it be $\\sigma\\colon K\\hookrightarrow \\mathbb{R}$, and</p>\n\n<p><strong>[H2]</strong> $K/\\mathbb{Q}$ is unramified at $2$.</p>\n\n<p>Consider then the following</p>\n\n<p><strong>Claim</strong> <em>An element in $\\mathcal{O}_K\\setminus\\{0\\}$ is a square if and only if it is a (global) norm from every quadratic extension $L/K$</em></p>\n\n<p>Admit the claim: we are going to deduce that if an element $x=r/s$ with $r,s\\in\\mathcal{O}_K$ is locally a square everywhere, then $x\\in(K^\\times)^2$. Let $y=xs^2$: then $y\\in\\mathcal{O}_K\\setminus\\{0\\}$ is locally a norm from every quadratic extension $L/K$, since it is a square in $K_\\mathfrak{p}^\\times$ for all $\\mathfrak{p}$. By the global norm theorem it is a global norm from every quadratic extension, and the claim tells you $y\\in (K^\\times)^2$: hence the same holds for $x=y/s^2$.</p>\n\n<p>To prove the claim, observe that one direction is obvious: squares are norms from every quadratic extension. On the other hand, let $x\\in\\mathcal{O}_K\\setminus\\{0\\}$ be a norm from every quadratic extension. I will use the real embedding $\\sigma$. Consider the function\n$$\nf\\colon \\mathcal{O}_K\\times\\mathcal{O}_K\\setminus \\{0\\}\\longrightarrow \\mathbb{R}\n$$\nsending $(a,b)$ to $f(a,b)=\\sigma\\left(\\left(\\frac{a}{b}\\right)^2-\\frac{x}{b^2}\\right)$. Then there is a bound $B\\in\\mathbb{R}$ such that $\\mathrm{Im}(f)\\subseteq [B,+\\infty)$. Indeed, $(a/b)^2\\geq 0$, while $b\\mapsto x/b^2$ is bounded from below, since $\\sigma(\\mathcal{O}_K)$ is discrete. Chose now a prime number $\\ell\\equiv 3\\pmod{4}$ which is unramfied in $K/\\mathbb{Q}$ and such that $-\\ell&lt;B$. Consider the extension $K(\\sqrt{-\\ell})/K$: since the discriminant of $\\mathcal{O}_K(\\sqrt{-\\ell})/\\mathcal{O}_K$ is $-4\\ell$ and the ideal $\\ell\\mathcal{O}_K$ is square-free by assumption, it follows that $[\\mathcal{O}_L:\\mathcal{O}_K(\\sqrt{-\\ell})]=1,2$. Since we assumed that $K/\\mathbb{Q}$ be unramified at $2$, the same argument as above implies that $\\mathcal{O}_{K(i)}=\\mathcal{O}_K[i]$ and then that $x^2+y^2\\equiv 0\\pmod{4\\mathcal{O}_K}$ has the unique solution $x\\equiv y\\equiv 0\\pmod{4\\mathcal{O}_K}$ (because each prime $\\mathfrak{r}\\mid 2$ of $\\mathcal{O}_K$ is ramified in $K(i)/K$, hence if something has norm from $K(i)$ to $K$ which is divisible by $4$, it must lie in $2\\mathcal{O}_K[i]$): in particular, the usual proof that the ring of integers of the imaginary quadratic field $\\mathbb{Q}(\\sqrt{-\\ell})$ is $\\mathbb{Z}[\\sqrt{-\\ell}]$ carries over, and $[\\mathcal{O}_L:\\mathcal{O}_K(\\sqrt{-\\ell})]=1$.</p>\n\n<p>Since we assumed that $x$ is a norm from $K(\\sqrt{-\\ell})/K$, we get $x=a^2+\\ell b^2$ for suitable $a,b\\in\\mathcal{O}_K$, by the above discussion. If $b=0$, then $x$ is a square, and we win. If not, rewrite the equality as\n$$\n\\frac{a^2-x}{b^2}=-\\ell\\Longrightarrow \\sigma(-\\ell)=-\\ell=f(a,b)\n$$\nThis contradicts the fact that $\\mathrm(Im)(f)\\subseteq [B,+\\infty)$ and finishes the proof of the claim, but I realise that one can reasonably object that the above proof is not \"quick\"...</p>\n\n<p>One final comment about assumption <strong>[H2]</strong>. It was only needed above to be sure that $\\mathcal{O}_{L}=\\mathcal{O}_K[\\sqrt{-\\ell}]$. This is not necessary: one could do, for instance, by proving that there must exist a $\\delta$ with $\\sigma(-\\delta)&lt;B$ and $\\mathcal{O}_{K(\\sqrt{-\\delta})}=\\mathcal{O}_K[\\sqrt{-\\delta}]$. Or, if one assumes that $2\\notin (K^\\times)^2$, then also the case $[\\mathcal{O}_{L}\\colon\\mathcal{O}_K]=2$ can be treated, by modifying a bit $f(a,b)$. But I could not do much better, hence I left assumption <strong>[H2]</strong>, although probably too restrective.</p>\n\n<p>Now, with global class field theory. The extension $L=K(\\sqrt{x})/K$ is abelian and corresponds to a unique normic subgroup $\\mathcal{H}\\subseteq C_K$ where $C_K$ is the idèle class group of $K$, so $\\mathcal{H}=\\mathrm{Norm}_{L/K}(C_L)$. Your assumption that $x\\in (K_\\mathfrak{p}^\\times)^2$ for all primes $\\mathfrak{p}$ means that $\\mathrm{Norm}_{L_\\mathfrak{P}/K_\\mathfrak{p}}(L_\\mathfrak{P}^\\times)=K_\\mathfrak{p}^\\times$ for all $\\mathfrak{p}$ and all $\\mathfrak{P}\\mid\\mathfrak{p}$ (<em>caution: we are using here that the minimal polynomial for $x$ has degree $1$ or $2$. Indeed, saying that $x\\in (K_\\mathfrak{p}^\\times)^2$ is equivalent to saying that all roots of $X^2-x$ are in $K_\\mathfrak{p}$, but this can be false if we replace $X^2-x$ by $X^n-x$ and try to prove the analogous statement replacing \"square\" by \"$n$th power\"</em>). Then $\\mathrm{Norm}_{L/K}(C_L)=C_K$ and hence $L=K$, or $x\\in(K^\\times)^2$. This is an adaptation of Corollary 8.8 in Tate's article on Global Class Field Theory in Cassels and Fröhlich's <em>Algebraic Number Theory</em>.</p>\n" } ]
Question: Does the Hasse norm theorem easily imply the global squares theorem? Answer 1: <p>I'm not sure how constructive the following answer to your second question is. If $\alpha$ is not a square, then there must exist a prime ideal ${\mathfrak p}$ coprime to $2\alpha$ with $(\alpha/{\mathfrak p}) = -1$. This follows from Kronecker's density theorem according to which the density of the prime ideals that split completely in a normal extension $L/K$ is equal to $1/(L:K)$. All you need for a proof is that the Dedekind zeta function of a number field $F$ has the property that $(s-1) \zeta_F(s)$ has a finite limit as $s \to 1$. </p> <p>Once you have such a prime ideal ${\mathfrak p}$, take any quadratic extension $L/K$ in which ${\mathfrak p}$ ramifies (you might pick an ideal ${\mathfrak a}$ in the inverse class, write ${\mathfrak p}{\mathfrak a} = (\beta)$, and set $L = K(\sqrt{\beta})$, for examples), and $\alpha$ will not be a norm from $L$.</p> <p>I would be suprised if you could construct ${\mathfrak p}$ "explicitly". Take e.g. the case where $\sqrt{\alpha}$ generates the Hilbert class field of a number field $K$ with class number $2$; then you would have to find a nonprincipal prime ideal in $K$.The fastest solution in this case is, as far as I know, picking prime ideals at random.</p> Answer 2: <p>Well, I would say that this crucially depends on what you define to be "quick". If you admit global class field theory, at least in its idelic formulation, the fact that all primes in $K$ are split in the extensions $K(\sqrt{x})/K$ proves that $x$ is a square quite rapidly (as Timo Keller has observed in the comments, this is actually easier than Chebotarev; and Franz Lemmermeyer's answer says even more). I give a short proof at the end of my answer, although it does not "need" the global norm theorem (which is anyhow hidden beneath the idelic formulation of global class field theory). I am pretty sure that this is the proof you have in mind when you ask for a <em>quicker one</em>.</p> <p>Without class field theory, I could come up only with an "elementary" proof assuming that </p> <p><strong>[H1]</strong> $K$ admits at least one real embedding, let it be $\sigma\colon K\hookrightarrow \mathbb{R}$, and</p> <p><strong>[H2]</strong> $K/\mathbb{Q}$ is unramified at $2$.</p> <p>Consider then the following</p> <p><strong>Claim</strong> <em>An element in $\mathcal{O}_K\setminus\{0\}$ is a square if and only if it is a (global) norm from every quadratic extension $L/K$</em></p> <p>Admit the claim: we are going to deduce that if an element $x=r/s$ with $r,s\in\mathcal{O}_K$ is locally a square everywhere, then $x\in(K^\times)^2$. Let $y=xs^2$: then $y\in\mathcal{O}_K\setminus\{0\}$ is locally a norm from every quadratic extension $L/K$, since it is a square in $K_\mathfrak{p}^\times$ for all $\mathfrak{p}$. By the global norm theorem it is a global norm from every quadratic extension, and the claim tells you $y\in (K^\times)^2$: hence the same holds for $x=y/s^2$.</p> <p>To prove the claim, observe that one direction is obvious: squares are norms from every quadratic extension. On the other hand, let $x\in\mathcal{O}_K\setminus\{0\}$ be a norm from every quadratic extension. I will use the real embedding $\sigma$. Consider the function $$ f\colon \mathcal{O}_K\times\mathcal{O}_K\setminus \{0\}\longrightarrow \mathbb{R} $$ sending $(a,b)$ to $f(a,b)=\sigma\left(\left(\frac{a}{b}\right)^2-\frac{x}{b^2}\right)$. Then there is a bound $B\in\mathbb{R}$ such that $\mathrm{Im}(f)\subseteq [B,+\infty)$. Indeed, $(a/b)^2\geq 0$, while $b\mapsto x/b^2$ is bounded from below, since $\sigma(\mathcal{O}_K)$ is discrete. Chose now a prime number $\ell\equiv 3\pmod{4}$ which is unramfied in $K/\mathbb{Q}$ and such that $-\ell&lt;B$. Consider the extension $K(\sqrt{-\ell})/K$: since the discriminant of $\mathcal{O}_K(\sqrt{-\ell})/\mathcal{O}_K$ is $-4\ell$ and the ideal $\ell\mathcal{O}_K$ is square-free by assumption, it follows that $[\mathcal{O}_L:\mathcal{O}_K(\sqrt{-\ell})]=1,2$. Since we assumed that $K/\mathbb{Q}$ be unramified at $2$, the same argument as above implies that $\mathcal{O}_{K(i)}=\mathcal{O}_K[i]$ and then that $x^2+y^2\equiv 0\pmod{4\mathcal{O}_K}$ has the unique solution $x\equiv y\equiv 0\pmod{4\mathcal{O}_K}$ (because each prime $\mathfrak{r}\mid 2$ of $\mathcal{O}_K$ is ramified in $K(i)/K$, hence if something has norm from $K(i)$ to $K$ which is divisible by $4$, it must lie in $2\mathcal{O}_K[i]$): in particular, the usual proof that the ring of integers of the imaginary quadratic field $\mathbb{Q}(\sqrt{-\ell})$ is $\mathbb{Z}[\sqrt{-\ell}]$ carries over, and $[\mathcal{O}_L:\mathcal{O}_K(\sqrt{-\ell})]=1$.</p> <p>Since we assumed that $x$ is a norm from $K(\sqrt{-\ell})/K$, we get $x=a^2+\ell b^2$ for suitable $a,b\in\mathcal{O}_K$, by the above discussion. If $b=0$, then $x$ is a square, and we win. If not, rewrite the equality as $$ \frac{a^2-x}{b^2}=-\ell\Longrightarrow \sigma(-\ell)=-\ell=f(a,b) $$ This contradicts the fact that $\mathrm(Im)(f)\subseteq [B,+\infty)$ and finishes the proof of the claim, but I realise that one can reasonably object that the above proof is not "quick"...</p> <p>One final comment about assumption <strong>[H2]</strong>. It was only needed above to be sure that $\mathcal{O}_{L}=\mathcal{O}_K[\sqrt{-\ell}]$. This is not necessary: one could do, for instance, by proving that there must exist a $\delta$ with $\sigma(-\delta)&lt;B$ and $\mathcal{O}_{K(\sqrt{-\delta})}=\mathcal{O}_K[\sqrt{-\delta}]$. Or, if one assumes that $2\notin (K^\times)^2$, then also the case $[\mathcal{O}_{L}\colon\mathcal{O}_K]=2$ can be treated, by modifying a bit $f(a,b)$. But I could not do much better, hence I left assumption <strong>[H2]</strong>, although probably too restrective.</p> <p>Now, with global class field theory. The extension $L=K(\sqrt{x})/K$ is abelian and corresponds to a unique normic subgroup $\mathcal{H}\subseteq C_K$ where $C_K$ is the idèle class group of $K$, so $\mathcal{H}=\mathrm{Norm}_{L/K}(C_L)$. Your assumption that $x\in (K_\mathfrak{p}^\times)^2$ for all primes $\mathfrak{p}$ means that $\mathrm{Norm}_{L_\mathfrak{P}/K_\mathfrak{p}}(L_\mathfrak{P}^\times)=K_\mathfrak{p}^\times$ for all $\mathfrak{p}$ and all $\mathfrak{P}\mid\mathfrak{p}$ (<em>caution: we are using here that the minimal polynomial for $x$ has degree $1$ or $2$. Indeed, saying that $x\in (K_\mathfrak{p}^\times)^2$ is equivalent to saying that all roots of $X^2-x$ are in $K_\mathfrak{p}$, but this can be false if we replace $X^2-x$ by $X^n-x$ and try to prove the analogous statement replacing "square" by "$n$th power"</em>). Then $\mathrm{Norm}_{L/K}(C_L)=C_K$ and hence $L=K$, or $x\in(K^\times)^2$. This is an adaptation of Corollary 8.8 in Tate's article on Global Class Field Theory in Cassels and Fröhlich's <em>Algebraic Number Theory</em>.</p> Answer 3: <p>I'm not sure how constructive the following answer to your second question is. If $\alpha$ is not a square, then there must exist a prime ideal ${\mathfrak p}$ coprime to $2\alpha$ with $(\alpha/{\mathfrak p}) = -1$. This follows from Kronecker's density theorem according to which the density of the prime ideals that split completely in a normal extension $L/K$ is equal to $1/(L:K)$. All you need for a proof is that the Dedekind zeta function of a number field $F$ has the property that $(s-1) \zeta_F(s)$ has a finite limit as $s \to 1$. </p> <p>Once you have such a prime ideal ${\mathfrak p}$, take any quadratic extension $L/K$ in which ${\mathfrak p}$ ramifies (you might pick an ideal ${\mathfrak a}$ in the inverse class, write ${\mathfrak p}{\mathfrak a} = (\beta)$, and set $L = K(\sqrt{\beta})$, for examples), and $\alpha$ will not be a norm from $L$.</p> <p>I would be suprised if you could construct ${\mathfrak p}$ "explicitly". Take e.g. the case where $\sqrt{\alpha}$ generates the Hilbert class field of a number field $K$ with class number $2$; then you would have to find a nonprincipal prime ideal in $K$.The fastest solution in this case is, as far as I know, picking prime ideals at random.</p> Answer 4: <p>Well, I would say that this crucially depends on what you define to be "quick". If you admit global class field theory, at least in its idelic formulation, the fact that all primes in $K$ are split in the extensions $K(\sqrt{x})/K$ proves that $x$ is a square quite rapidly (as Timo Keller has observed in the comments, this is actually easier than Chebotarev; and Franz Lemmermeyer's answer says even more). I give a short proof at the end of my answer, although it does not "need" the global norm theorem (which is anyhow hidden beneath the idelic formulation of global class field theory). I am pretty sure that this is the proof you have in mind when you ask for a <em>quicker one</em>.</p> <p>Without class field theory, I could come up only with an "elementary" proof assuming that </p> <p><strong>[H1]</strong> $K$ admits at least one real embedding, let it be $\sigma\colon K\hookrightarrow \mathbb{R}$, and</p> <p><strong>[H2]</strong> $K/\mathbb{Q}$ is unramified at $2$.</p> <p>Consider then the following</p> <p><strong>Claim</strong> <em>An element in $\mathcal{O}_K\setminus\{0\}$ is a square if and only if it is a (global) norm from every quadratic extension $L/K$</em></p> <p>Admit the claim: we are going to deduce that if an element $x=r/s$ with $r,s\in\mathcal{O}_K$ is locally a square everywhere, then $x\in(K^\times)^2$. Let $y=xs^2$: then $y\in\mathcal{O}_K\setminus\{0\}$ is locally a norm from every quadratic extension $L/K$, since it is a square in $K_\mathfrak{p}^\times$ for all $\mathfrak{p}$. By the global norm theorem it is a global norm from every quadratic extension, and the claim tells you $y\in (K^\times)^2$: hence the same holds for $x=y/s^2$.</p> <p>To prove the claim, observe that one direction is obvious: squares are norms from every quadratic extension. On the other hand, let $x\in\mathcal{O}_K\setminus\{0\}$ be a norm from every quadratic extension. I will use the real embedding $\sigma$. Consider the function $$ f\colon \mathcal{O}_K\times\mathcal{O}_K\setminus \{0\}\longrightarrow \mathbb{R} $$ sending $(a,b)$ to $f(a,b)=\sigma\left(\left(\frac{a}{b}\right)^2-\frac{x}{b^2}\right)$. Then there is a bound $B\in\mathbb{R}$ such that $\mathrm{Im}(f)\subseteq [B,+\infty)$. Indeed, $(a/b)^2\geq 0$, while $b\mapsto x/b^2$ is bounded from below, since $\sigma(\mathcal{O}_K)$ is discrete. Chose now a prime number $\ell\equiv 3\pmod{4}$ which is unramfied in $K/\mathbb{Q}$ and such that $-\ell&lt;B$. Consider the extension $K(\sqrt{-\ell})/K$: since the discriminant of $\mathcal{O}_K(\sqrt{-\ell})/\mathcal{O}_K$ is $-4\ell$ and the ideal $\ell\mathcal{O}_K$ is square-free by assumption, it follows that $[\mathcal{O}_L:\mathcal{O}_K(\sqrt{-\ell})]=1,2$. Since we assumed that $K/\mathbb{Q}$ be unramified at $2$, the same argument as above implies that $\mathcal{O}_{K(i)}=\mathcal{O}_K[i]$ and then that $x^2+y^2\equiv 0\pmod{4\mathcal{O}_K}$ has the unique solution $x\equiv y\equiv 0\pmod{4\mathcal{O}_K}$ (because each prime $\mathfrak{r}\mid 2$ of $\mathcal{O}_K$ is ramified in $K(i)/K$, hence if something has norm from $K(i)$ to $K$ which is divisible by $4$, it must lie in $2\mathcal{O}_K[i]$): in particular, the usual proof that the ring of integers of the imaginary quadratic field $\mathbb{Q}(\sqrt{-\ell})$ is $\mathbb{Z}[\sqrt{-\ell}]$ carries over, and $[\mathcal{O}_L:\mathcal{O}_K(\sqrt{-\ell})]=1$.</p> <p>Since we assumed that $x$ is a norm from $K(\sqrt{-\ell})/K$, we get $x=a^2+\ell b^2$ for suitable $a,b\in\mathcal{O}_K$, by the above discussion. If $b=0$, then $x$ is a square, and we win. If not, rewrite the equality as $$ \frac{a^2-x}{b^2}=-\ell\Longrightarrow \sigma(-\ell)=-\ell=f(a,b) $$ This contradicts the fact that $\mathrm(Im)(f)\subseteq [B,+\infty)$ and finishes the proof of the claim, but I realise that one can reasonably object that the above proof is not "quick"...</p> <p>One final comment about assumption <strong>[H2]</strong>. It was only needed above to be sure that $\mathcal{O}_{L}=\mathcal{O}_K[\sqrt{-\ell}]$. This is not necessary: one could do, for instance, by proving that there must exist a $\delta$ with $\sigma(-\delta)&lt;B$ and $\mathcal{O}_{K(\sqrt{-\delta})}=\mathcal{O}_K[\sqrt{-\delta}]$. Or, if one assumes that $2\notin (K^\times)^2$, then also the case $[\mathcal{O}_{L}\colon\mathcal{O}_K]=2$ can be treated, by modifying a bit $f(a,b)$. But I could not do much better, hence I left assumption <strong>[H2]</strong>, although probably too restrective.</p> <p>Now, with global class field theory. The extension $L=K(\sqrt{x})/K$ is abelian and corresponds to a unique normic subgroup $\mathcal{H}\subseteq C_K$ where $C_K$ is the idèle class group of $K$, so $\mathcal{H}=\mathrm{Norm}_{L/K}(C_L)$. Your assumption that $x\in (K_\mathfrak{p}^\times)^2$ for all primes $\mathfrak{p}$ means that $\mathrm{Norm}_{L_\mathfrak{P}/K_\mathfrak{p}}(L_\mathfrak{P}^\times)=K_\mathfrak{p}^\times$ for all $\mathfrak{p}$ and all $\mathfrak{P}\mid\mathfrak{p}$ (<em>caution: we are using here that the minimal polynomial for $x$ has degree $1$ or $2$. Indeed, saying that $x\in (K_\mathfrak{p}^\times)^2$ is equivalent to saying that all roots of $X^2-x$ are in $K_\mathfrak{p}$, but this can be false if we replace $X^2-x$ by $X^n-x$ and try to prove the analogous statement replacing "square" by "$n$th power"</em>). Then $\mathrm{Norm}_{L/K}(C_L)=C_K$ and hence $L=K$, or $x\in(K^\times)^2$. This is an adaptation of Corollary 8.8 in Tate's article on Global Class Field Theory in Cassels and Fröhlich's <em>Algebraic Number Theory</em>.</p> Answer 5: <p>I'm not sure how constructive the following answer to your second question is. If $\alpha$ is not a square, then there must exist a prime ideal ${\mathfrak p}$ coprime to $2\alpha$ with $(\alpha/{\mathfrak p}) = -1$. This follows from Kronecker's density theorem according to which the density of the prime ideals that split completely in a normal extension $L/K$ is equal to $1/(L:K)$. All you need for a proof is that the Dedekind zeta function of a number field $F$ has the property that $(s-1) \zeta_F(s)$ has a finite limit as $s \to 1$. </p> <p>Once you have such a prime ideal ${\mathfrak p}$, take any quadratic extension $L/K$ in which ${\mathfrak p}$ ramifies (you might pick an ideal ${\mathfrak a}$ in the inverse class, write ${\mathfrak p}{\mathfrak a} = (\beta)$, and set $L = K(\sqrt{\beta})$, for examples), and $\alpha$ will not be a norm from $L$.</p> <p>I would be suprised if you could construct ${\mathfrak p}$ "explicitly". Take e.g. the case where $\sqrt{\alpha}$ generates the Hilbert class field of a number field $K$ with class number $2$; then you would have to find a nonprincipal prime ideal in $K$.The fastest solution in this case is, as far as I know, picking prime ideals at random.</p> Answer 6: <p>Well, I would say that this crucially depends on what you define to be "quick". If you admit global class field theory, at least in its idelic formulation, the fact that all primes in $K$ are split in the extensions $K(\sqrt{x})/K$ proves that $x$ is a square quite rapidly (as Timo Keller has observed in the comments, this is actually easier than Chebotarev; and Franz Lemmermeyer's answer says even more). I give a short proof at the end of my answer, although it does not "need" the global norm theorem (which is anyhow hidden beneath the idelic formulation of global class field theory). I am pretty sure that this is the proof you have in mind when you ask for a <em>quicker one</em>.</p> <p>Without class field theory, I could come up only with an "elementary" proof assuming that </p> <p><strong>[H1]</strong> $K$ admits at least one real embedding, let it be $\sigma\colon K\hookrightarrow \mathbb{R}$, and</p> <p><strong>[H2]</strong> $K/\mathbb{Q}$ is unramified at $2$.</p> <p>Consider then the following</p> <p><strong>Claim</strong> <em>An element in $\mathcal{O}_K\setminus\{0\}$ is a square if and only if it is a (global) norm from every quadratic extension $L/K$</em></p> <p>Admit the claim: we are going to deduce that if an element $x=r/s$ with $r,s\in\mathcal{O}_K$ is locally a square everywhere, then $x\in(K^\times)^2$. Let $y=xs^2$: then $y\in\mathcal{O}_K\setminus\{0\}$ is locally a norm from every quadratic extension $L/K$, since it is a square in $K_\mathfrak{p}^\times$ for all $\mathfrak{p}$. By the global norm theorem it is a global norm from every quadratic extension, and the claim tells you $y\in (K^\times)^2$: hence the same holds for $x=y/s^2$.</p> <p>To prove the claim, observe that one direction is obvious: squares are norms from every quadratic extension. On the other hand, let $x\in\mathcal{O}_K\setminus\{0\}$ be a norm from every quadratic extension. I will use the real embedding $\sigma$. Consider the function $$ f\colon \mathcal{O}_K\times\mathcal{O}_K\setminus \{0\}\longrightarrow \mathbb{R} $$ sending $(a,b)$ to $f(a,b)=\sigma\left(\left(\frac{a}{b}\right)^2-\frac{x}{b^2}\right)$. Then there is a bound $B\in\mathbb{R}$ such that $\mathrm{Im}(f)\subseteq [B,+\infty)$. Indeed, $(a/b)^2\geq 0$, while $b\mapsto x/b^2$ is bounded from below, since $\sigma(\mathcal{O}_K)$ is discrete. Chose now a prime number $\ell\equiv 3\pmod{4}$ which is unramfied in $K/\mathbb{Q}$ and such that $-\ell&lt;B$. Consider the extension $K(\sqrt{-\ell})/K$: since the discriminant of $\mathcal{O}_K(\sqrt{-\ell})/\mathcal{O}_K$ is $-4\ell$ and the ideal $\ell\mathcal{O}_K$ is square-free by assumption, it follows that $[\mathcal{O}_L:\mathcal{O}_K(\sqrt{-\ell})]=1,2$. Since we assumed that $K/\mathbb{Q}$ be unramified at $2$, the same argument as above implies that $\mathcal{O}_{K(i)}=\mathcal{O}_K[i]$ and then that $x^2+y^2\equiv 0\pmod{4\mathcal{O}_K}$ has the unique solution $x\equiv y\equiv 0\pmod{4\mathcal{O}_K}$ (because each prime $\mathfrak{r}\mid 2$ of $\mathcal{O}_K$ is ramified in $K(i)/K$, hence if something has norm from $K(i)$ to $K$ which is divisible by $4$, it must lie in $2\mathcal{O}_K[i]$): in particular, the usual proof that the ring of integers of the imaginary quadratic field $\mathbb{Q}(\sqrt{-\ell})$ is $\mathbb{Z}[\sqrt{-\ell}]$ carries over, and $[\mathcal{O}_L:\mathcal{O}_K(\sqrt{-\ell})]=1$.</p> <p>Since we assumed that $x$ is a norm from $K(\sqrt{-\ell})/K$, we get $x=a^2+\ell b^2$ for suitable $a,b\in\mathcal{O}_K$, by the above discussion. If $b=0$, then $x$ is a square, and we win. If not, rewrite the equality as $$ \frac{a^2-x}{b^2}=-\ell\Longrightarrow \sigma(-\ell)=-\ell=f(a,b) $$ This contradicts the fact that $\mathrm(Im)(f)\subseteq [B,+\infty)$ and finishes the proof of the claim, but I realise that one can reasonably object that the above proof is not "quick"...</p> <p>One final comment about assumption <strong>[H2]</strong>. It was only needed above to be sure that $\mathcal{O}_{L}=\mathcal{O}_K[\sqrt{-\ell}]$. This is not necessary: one could do, for instance, by proving that there must exist a $\delta$ with $\sigma(-\delta)&lt;B$ and $\mathcal{O}_{K(\sqrt{-\delta})}=\mathcal{O}_K[\sqrt{-\delta}]$. Or, if one assumes that $2\notin (K^\times)^2$, then also the case $[\mathcal{O}_{L}\colon\mathcal{O}_K]=2$ can be treated, by modifying a bit $f(a,b)$. But I could not do much better, hence I left assumption <strong>[H2]</strong>, although probably too restrective.</p> <p>Now, with global class field theory. The extension $L=K(\sqrt{x})/K$ is abelian and corresponds to a unique normic subgroup $\mathcal{H}\subseteq C_K$ where $C_K$ is the idèle class group of $K$, so $\mathcal{H}=\mathrm{Norm}_{L/K}(C_L)$. Your assumption that $x\in (K_\mathfrak{p}^\times)^2$ for all primes $\mathfrak{p}$ means that $\mathrm{Norm}_{L_\mathfrak{P}/K_\mathfrak{p}}(L_\mathfrak{P}^\times)=K_\mathfrak{p}^\times$ for all $\mathfrak{p}$ and all $\mathfrak{P}\mid\mathfrak{p}$ (<em>caution: we are using here that the minimal polynomial for $x$ has degree $1$ or $2$. Indeed, saying that $x\in (K_\mathfrak{p}^\times)^2$ is equivalent to saying that all roots of $X^2-x$ are in $K_\mathfrak{p}$, but this can be false if we replace $X^2-x$ by $X^n-x$ and try to prove the analogous statement replacing "square" by "$n$th power"</em>). Then $\mathrm{Norm}_{L/K}(C_L)=C_K$ and hence $L=K$, or $x\in(K^\times)^2$. This is an adaptation of Corollary 8.8 in Tate's article on Global Class Field Theory in Cassels and Fröhlich's <em>Algebraic Number Theory</em>.</p>
Most important mathematical results in last 30 years
259,804
10
https://mathoverflow.net/questions/259804/most-important-mathematical-results-in-last-30-years
<p>Which results from the last 30 years, in any area of mathematics, do you think are the most important ones?</p> <p>Specifically, which are the ones that will have more impact across all math and/or settle significant questions, in your opinion? It does not matter if they are big, lond-proofed results like a classification theorem, or little, elementary lemmas proved in a paragraph.</p> <p>To keep this reasonable, I propose that every contributor posts at most 10 results.</p> <p>Please, write one result per post, so that we can all vote properly and have a real ranking afterwards. Also, please don't just write the name, write a little description (hopefully, self-contained) and put some link if desirable.</p> <p>(My personal aim with this question is to learn. I hope to be a bit better versed after studying your answers!)</p>
[ { "answer_id": 259811, "score": 11, "text": "<p>Taniyama, Shimura and later Weil conjectured around 60 years ago that the L-function of an elliptic curve arises from a modular form. This conjecture was known to entail the last Fermat theorem after pioneering works of Hellegouarch, Frey and Ribet. In 1994, after a very long and technical work made in isolation and relying on Galois representations, Wiles managed to prove this conjecture in the semi-stable case, which was enough to establish the veracity of the now called Fermat-Wiles theorem. Based on this work, an article published in 1999 by Breuil, Conrad, Diamond and Taylor establish the result in full generality.</p>\n\n<p>This theorem is a tiny part of the Langlands program, dating from 1967.</p>\n" }, { "answer_id": 259805, "score": 10, "text": "<p>The proof of the <strong>Milnor-Bloch-Kato conjecture</strong> by <em>Vladimir Voevodsky</em> (Fields medal 2002):</p>\n\n<p>One has a norm residue isomorphism $$\\mathrm{K}^\\mathrm{M}_n(k)/\\ell \\to \\mathrm{H}^n(k, \\mu_\\ell^{\\otimes n})$$ from Milnor $K$-theory to étale cohomology, $\\ell$ invertible in $k$.</p>\n\n<p>This is translated into a statement of the comparison of the cohomology of motivic complexes with respect to different topologies.</p>\n\n<p>For $\\ell = 2$, this is the Milnor conjecture. Even for this, Voevodsky got the Fields medal.</p>\n\n<p>For $n = 0$, this is obvious, for $n = 1$, it follows from Hilbert's Theorem 90. A corollary is that the Galois cohomology ring is generated by elements of degree $1$.</p>\n\n<p>It implies the Beilinson-Lichtenbaum conjecture, a more general conjecture for varieties, not just fields.</p>\n" }, { "answer_id": 259807, "score": 6, "text": "<p>In discrete math:</p>\n\n<p>Keevash's probabilisitic proof of Steiner's conjecture (On the Existence of Designs) is perhaps the biggest recent-ish result in combinatorics [due to the problem's ridiculously long pedigree].</p>\n\n<p>But if you allow a little further back, I think Szemeredi's regularity lemma would have to take the cake for its outrageous applicability.</p>\n" }, { "answer_id": 259809, "score": 6, "text": "<p>Probably one of the major developments in homotopy theory is the <a href=\"https://arxiv.org/abs/0908.3724\">solution to the Kervaire invariant one problem</a> by Hill, Hopkins and Ravenel. It showcased the power of equivariant homotopy theory in solving a problem that a priori has nothing to do with it. In particular the idea to use the slice spectral sequence to get a handle on the homotopy fixed point spectral sequence of a certain spectrum is a real eye-opener.</p>\n" }, { "answer_id": 259926, "score": 6, "text": "<p>Perelman's proof of the Geometrization conjecture (see <a href=\"https://arxiv.org/abs/math/0211159\" rel=\"nofollow noreferrer\">here</a>, <a href=\"https://arxiv.org/abs/math/0303109\" rel=\"nofollow noreferrer\">here</a> and <a href=\"https://arxiv.org/abs/math/0307245\" rel=\"nofollow noreferrer\">here</a>) was the crowning achievement of decades of work. It was the most important of Thurston's conjectures about the topology of 3-manifolds.</p>\n\n<p>Oh, and I nearly forgot to mention that the Poincare Conjecture is a consequence.</p>\n" }, { "answer_id": 259925, "score": 5, "text": "<p>Agol's proof of the <a href=\"http://www.math.uiuc.edu/documenta/vol-18/vol-18-eng.html\" rel=\"noreferrer\">Virtual Haken conjecture</a> was a wonderful application of the tools developed by Wise and his coauthors in geometric group theory to 3-manifold topology. The Virtual Haken conjecture, which can be thought of as the topological classification of compact 3-manifolds, is a fundamental result.</p>\n\n<p>Agol's theorem:</p>\n\n<blockquote>\n <p>Every cubulable hyperbolic group is virtually special.</p>\n</blockquote>\n\n<p>also has other dramatic consequences for hyperbolic groups. It's arguably the most important theorem proved in both algebra <em>and</em> topology in the last five years.</p>\n" }, { "answer_id": 259927, "score": 3, "text": "<p>Kahn--Markovic's proofs of the <a href=\"https://arxiv.org/abs/0910.5501\" rel=\"nofollow noreferrer\">Surface Subgroup conjecture</a> and the <a href=\"https://arxiv.org/abs/1101.1330\" rel=\"nofollow noreferrer\">Ehrenpreis conjecture</a>.</p>\n" }, { "answer_id": 259806, "score": 2, "text": "<p>I'm going for:</p>\n\n<blockquote>\n <p><a href=\"http://mathworld.wolfram.com/GausssClassNumberProblem.html\" rel=\"nofollow noreferrer\">Gauss's Class Number Problem</a></p>\n</blockquote>\n\n<p>solved for $m\\le100$ by Watkins in 2004, in '<a href=\"http://www.ams.org/journals/mcom/2004-73-246/S0025-5718-03-01517-5/S0025-5718-03-01517-5.pdf\" rel=\"nofollow noreferrer\">Class Numbers of Imaginary Quadratic Fields</a>'.</p>\n" }, { "answer_id": 259811, "score": 11, "text": "<p>Taniyama, Shimura and later Weil conjectured around 60 years ago that the L-function of an elliptic curve arises from a modular form. This conjecture was known to entail the last Fermat theorem after pioneering works of Hellegouarch, Frey and Ribet. In 1994, after a very long and technical work made in isolation and relying on Galois representations, Wiles managed to prove this conjecture in the semi-stable case, which was enough to establish the veracity of the now called Fermat-Wiles theorem. Based on this work, an article published in 1999 by Breuil, Conrad, Diamond and Taylor establish the result in full generality.</p>\n\n<p>This theorem is a tiny part of the Langlands program, dating from 1967.</p>\n" }, { "answer_id": 259805, "score": 10, "text": "<p>The proof of the <strong>Milnor-Bloch-Kato conjecture</strong> by <em>Vladimir Voevodsky</em> (Fields medal 2002):</p>\n\n<p>One has a norm residue isomorphism $$\\mathrm{K}^\\mathrm{M}_n(k)/\\ell \\to \\mathrm{H}^n(k, \\mu_\\ell^{\\otimes n})$$ from Milnor $K$-theory to étale cohomology, $\\ell$ invertible in $k$.</p>\n\n<p>This is translated into a statement of the comparison of the cohomology of motivic complexes with respect to different topologies.</p>\n\n<p>For $\\ell = 2$, this is the Milnor conjecture. Even for this, Voevodsky got the Fields medal.</p>\n\n<p>For $n = 0$, this is obvious, for $n = 1$, it follows from Hilbert's Theorem 90. A corollary is that the Galois cohomology ring is generated by elements of degree $1$.</p>\n\n<p>It implies the Beilinson-Lichtenbaum conjecture, a more general conjecture for varieties, not just fields.</p>\n" }, { "answer_id": 259807, "score": 6, "text": "<p>In discrete math:</p>\n\n<p>Keevash's probabilisitic proof of Steiner's conjecture (On the Existence of Designs) is perhaps the biggest recent-ish result in combinatorics [due to the problem's ridiculously long pedigree].</p>\n\n<p>But if you allow a little further back, I think Szemeredi's regularity lemma would have to take the cake for its outrageous applicability.</p>\n" }, { "answer_id": 259809, "score": 6, "text": "<p>Probably one of the major developments in homotopy theory is the <a href=\"https://arxiv.org/abs/0908.3724\">solution to the Kervaire invariant one problem</a> by Hill, Hopkins and Ravenel. It showcased the power of equivariant homotopy theory in solving a problem that a priori has nothing to do with it. In particular the idea to use the slice spectral sequence to get a handle on the homotopy fixed point spectral sequence of a certain spectrum is a real eye-opener.</p>\n" }, { "answer_id": 259926, "score": 6, "text": "<p>Perelman's proof of the Geometrization conjecture (see <a href=\"https://arxiv.org/abs/math/0211159\" rel=\"nofollow noreferrer\">here</a>, <a href=\"https://arxiv.org/abs/math/0303109\" rel=\"nofollow noreferrer\">here</a> and <a href=\"https://arxiv.org/abs/math/0307245\" rel=\"nofollow noreferrer\">here</a>) was the crowning achievement of decades of work. It was the most important of Thurston's conjectures about the topology of 3-manifolds.</p>\n\n<p>Oh, and I nearly forgot to mention that the Poincare Conjecture is a consequence.</p>\n" }, { "answer_id": 259925, "score": 5, "text": "<p>Agol's proof of the <a href=\"http://www.math.uiuc.edu/documenta/vol-18/vol-18-eng.html\" rel=\"noreferrer\">Virtual Haken conjecture</a> was a wonderful application of the tools developed by Wise and his coauthors in geometric group theory to 3-manifold topology. The Virtual Haken conjecture, which can be thought of as the topological classification of compact 3-manifolds, is a fundamental result.</p>\n\n<p>Agol's theorem:</p>\n\n<blockquote>\n <p>Every cubulable hyperbolic group is virtually special.</p>\n</blockquote>\n\n<p>also has other dramatic consequences for hyperbolic groups. It's arguably the most important theorem proved in both algebra <em>and</em> topology in the last five years.</p>\n" }, { "answer_id": 259927, "score": 3, "text": "<p>Kahn--Markovic's proofs of the <a href=\"https://arxiv.org/abs/0910.5501\" rel=\"nofollow noreferrer\">Surface Subgroup conjecture</a> and the <a href=\"https://arxiv.org/abs/1101.1330\" rel=\"nofollow noreferrer\">Ehrenpreis conjecture</a>.</p>\n" }, { "answer_id": 259806, "score": 2, "text": "<p>I'm going for:</p>\n\n<blockquote>\n <p><a href=\"http://mathworld.wolfram.com/GausssClassNumberProblem.html\" rel=\"nofollow noreferrer\">Gauss's Class Number Problem</a></p>\n</blockquote>\n\n<p>solved for $m\\le100$ by Watkins in 2004, in '<a href=\"http://www.ams.org/journals/mcom/2004-73-246/S0025-5718-03-01517-5/S0025-5718-03-01517-5.pdf\" rel=\"nofollow noreferrer\">Class Numbers of Imaginary Quadratic Fields</a>'.</p>\n" }, { "answer_id": 259811, "score": 11, "text": "<p>Taniyama, Shimura and later Weil conjectured around 60 years ago that the L-function of an elliptic curve arises from a modular form. This conjecture was known to entail the last Fermat theorem after pioneering works of Hellegouarch, Frey and Ribet. In 1994, after a very long and technical work made in isolation and relying on Galois representations, Wiles managed to prove this conjecture in the semi-stable case, which was enough to establish the veracity of the now called Fermat-Wiles theorem. Based on this work, an article published in 1999 by Breuil, Conrad, Diamond and Taylor establish the result in full generality.</p>\n\n<p>This theorem is a tiny part of the Langlands program, dating from 1967.</p>\n" }, { "answer_id": 259805, "score": 10, "text": "<p>The proof of the <strong>Milnor-Bloch-Kato conjecture</strong> by <em>Vladimir Voevodsky</em> (Fields medal 2002):</p>\n\n<p>One has a norm residue isomorphism $$\\mathrm{K}^\\mathrm{M}_n(k)/\\ell \\to \\mathrm{H}^n(k, \\mu_\\ell^{\\otimes n})$$ from Milnor $K$-theory to étale cohomology, $\\ell$ invertible in $k$.</p>\n\n<p>This is translated into a statement of the comparison of the cohomology of motivic complexes with respect to different topologies.</p>\n\n<p>For $\\ell = 2$, this is the Milnor conjecture. Even for this, Voevodsky got the Fields medal.</p>\n\n<p>For $n = 0$, this is obvious, for $n = 1$, it follows from Hilbert's Theorem 90. A corollary is that the Galois cohomology ring is generated by elements of degree $1$.</p>\n\n<p>It implies the Beilinson-Lichtenbaum conjecture, a more general conjecture for varieties, not just fields.</p>\n" }, { "answer_id": 259807, "score": 6, "text": "<p>In discrete math:</p>\n\n<p>Keevash's probabilisitic proof of Steiner's conjecture (On the Existence of Designs) is perhaps the biggest recent-ish result in combinatorics [due to the problem's ridiculously long pedigree].</p>\n\n<p>But if you allow a little further back, I think Szemeredi's regularity lemma would have to take the cake for its outrageous applicability.</p>\n" }, { "answer_id": 259809, "score": 6, "text": "<p>Probably one of the major developments in homotopy theory is the <a href=\"https://arxiv.org/abs/0908.3724\">solution to the Kervaire invariant one problem</a> by Hill, Hopkins and Ravenel. It showcased the power of equivariant homotopy theory in solving a problem that a priori has nothing to do with it. In particular the idea to use the slice spectral sequence to get a handle on the homotopy fixed point spectral sequence of a certain spectrum is a real eye-opener.</p>\n" }, { "answer_id": 259926, "score": 6, "text": "<p>Perelman's proof of the Geometrization conjecture (see <a href=\"https://arxiv.org/abs/math/0211159\" rel=\"nofollow noreferrer\">here</a>, <a href=\"https://arxiv.org/abs/math/0303109\" rel=\"nofollow noreferrer\">here</a> and <a href=\"https://arxiv.org/abs/math/0307245\" rel=\"nofollow noreferrer\">here</a>) was the crowning achievement of decades of work. It was the most important of Thurston's conjectures about the topology of 3-manifolds.</p>\n\n<p>Oh, and I nearly forgot to mention that the Poincare Conjecture is a consequence.</p>\n" }, { "answer_id": 259925, "score": 5, "text": "<p>Agol's proof of the <a href=\"http://www.math.uiuc.edu/documenta/vol-18/vol-18-eng.html\" rel=\"noreferrer\">Virtual Haken conjecture</a> was a wonderful application of the tools developed by Wise and his coauthors in geometric group theory to 3-manifold topology. The Virtual Haken conjecture, which can be thought of as the topological classification of compact 3-manifolds, is a fundamental result.</p>\n\n<p>Agol's theorem:</p>\n\n<blockquote>\n <p>Every cubulable hyperbolic group is virtually special.</p>\n</blockquote>\n\n<p>also has other dramatic consequences for hyperbolic groups. It's arguably the most important theorem proved in both algebra <em>and</em> topology in the last five years.</p>\n" }, { "answer_id": 259927, "score": 3, "text": "<p>Kahn--Markovic's proofs of the <a href=\"https://arxiv.org/abs/0910.5501\" rel=\"nofollow noreferrer\">Surface Subgroup conjecture</a> and the <a href=\"https://arxiv.org/abs/1101.1330\" rel=\"nofollow noreferrer\">Ehrenpreis conjecture</a>.</p>\n" }, { "answer_id": 259806, "score": 2, "text": "<p>I'm going for:</p>\n\n<blockquote>\n <p><a href=\"http://mathworld.wolfram.com/GausssClassNumberProblem.html\" rel=\"nofollow noreferrer\">Gauss's Class Number Problem</a></p>\n</blockquote>\n\n<p>solved for $m\\le100$ by Watkins in 2004, in '<a href=\"http://www.ams.org/journals/mcom/2004-73-246/S0025-5718-03-01517-5/S0025-5718-03-01517-5.pdf\" rel=\"nofollow noreferrer\">Class Numbers of Imaginary Quadratic Fields</a>'.</p>\n" } ]
Question: Most important mathematical results in last 30 years Answer 1: <p>Taniyama, Shimura and later Weil conjectured around 60 years ago that the L-function of an elliptic curve arises from a modular form. This conjecture was known to entail the last Fermat theorem after pioneering works of Hellegouarch, Frey and Ribet. In 1994, after a very long and technical work made in isolation and relying on Galois representations, Wiles managed to prove this conjecture in the semi-stable case, which was enough to establish the veracity of the now called Fermat-Wiles theorem. Based on this work, an article published in 1999 by Breuil, Conrad, Diamond and Taylor establish the result in full generality.</p> <p>This theorem is a tiny part of the Langlands program, dating from 1967.</p> Answer 2: <p>The proof of the <strong>Milnor-Bloch-Kato conjecture</strong> by <em>Vladimir Voevodsky</em> (Fields medal 2002):</p> <p>One has a norm residue isomorphism $$\mathrm{K}^\mathrm{M}_n(k)/\ell \to \mathrm{H}^n(k, \mu_\ell^{\otimes n})$$ from Milnor $K$-theory to étale cohomology, $\ell$ invertible in $k$.</p> <p>This is translated into a statement of the comparison of the cohomology of motivic complexes with respect to different topologies.</p> <p>For $\ell = 2$, this is the Milnor conjecture. Even for this, Voevodsky got the Fields medal.</p> <p>For $n = 0$, this is obvious, for $n = 1$, it follows from Hilbert's Theorem 90. A corollary is that the Galois cohomology ring is generated by elements of degree $1$.</p> <p>It implies the Beilinson-Lichtenbaum conjecture, a more general conjecture for varieties, not just fields.</p> Answer 3: <p>In discrete math:</p> <p>Keevash's probabilisitic proof of Steiner's conjecture (On the Existence of Designs) is perhaps the biggest recent-ish result in combinatorics [due to the problem's ridiculously long pedigree].</p> <p>But if you allow a little further back, I think Szemeredi's regularity lemma would have to take the cake for its outrageous applicability.</p> Answer 4: <p>Probably one of the major developments in homotopy theory is the <a href="https://arxiv.org/abs/0908.3724">solution to the Kervaire invariant one problem</a> by Hill, Hopkins and Ravenel. It showcased the power of equivariant homotopy theory in solving a problem that a priori has nothing to do with it. In particular the idea to use the slice spectral sequence to get a handle on the homotopy fixed point spectral sequence of a certain spectrum is a real eye-opener.</p> Answer 5: <p>Perelman's proof of the Geometrization conjecture (see <a href="https://arxiv.org/abs/math/0211159" rel="nofollow noreferrer">here</a>, <a href="https://arxiv.org/abs/math/0303109" rel="nofollow noreferrer">here</a> and <a href="https://arxiv.org/abs/math/0307245" rel="nofollow noreferrer">here</a>) was the crowning achievement of decades of work. It was the most important of Thurston's conjectures about the topology of 3-manifolds.</p> <p>Oh, and I nearly forgot to mention that the Poincare Conjecture is a consequence.</p> Answer 6: <p>Agol's proof of the <a href="http://www.math.uiuc.edu/documenta/vol-18/vol-18-eng.html" rel="noreferrer">Virtual Haken conjecture</a> was a wonderful application of the tools developed by Wise and his coauthors in geometric group theory to 3-manifold topology. The Virtual Haken conjecture, which can be thought of as the topological classification of compact 3-manifolds, is a fundamental result.</p> <p>Agol's theorem:</p> <blockquote> <p>Every cubulable hyperbolic group is virtually special.</p> </blockquote> <p>also has other dramatic consequences for hyperbolic groups. It's arguably the most important theorem proved in both algebra <em>and</em> topology in the last five years.</p> Answer 7: <p>Kahn--Markovic's proofs of the <a href="https://arxiv.org/abs/0910.5501" rel="nofollow noreferrer">Surface Subgroup conjecture</a> and the <a href="https://arxiv.org/abs/1101.1330" rel="nofollow noreferrer">Ehrenpreis conjecture</a>.</p> Answer 8: <p>I'm going for:</p> <blockquote> <p><a href="http://mathworld.wolfram.com/GausssClassNumberProblem.html" rel="nofollow noreferrer">Gauss's Class Number Problem</a></p> </blockquote> <p>solved for $m\le100$ by Watkins in 2004, in '<a href="http://www.ams.org/journals/mcom/2004-73-246/S0025-5718-03-01517-5/S0025-5718-03-01517-5.pdf" rel="nofollow noreferrer">Class Numbers of Imaginary Quadratic Fields</a>'.</p> Answer 9: <p>Taniyama, Shimura and later Weil conjectured around 60 years ago that the L-function of an elliptic curve arises from a modular form. This conjecture was known to entail the last Fermat theorem after pioneering works of Hellegouarch, Frey and Ribet. In 1994, after a very long and technical work made in isolation and relying on Galois representations, Wiles managed to prove this conjecture in the semi-stable case, which was enough to establish the veracity of the now called Fermat-Wiles theorem. Based on this work, an article published in 1999 by Breuil, Conrad, Diamond and Taylor establish the result in full generality.</p> <p>This theorem is a tiny part of the Langlands program, dating from 1967.</p> Answer 10: <p>The proof of the <strong>Milnor-Bloch-Kato conjecture</strong> by <em>Vladimir Voevodsky</em> (Fields medal 2002):</p> <p>One has a norm residue isomorphism $$\mathrm{K}^\mathrm{M}_n(k)/\ell \to \mathrm{H}^n(k, \mu_\ell^{\otimes n})$$ from Milnor $K$-theory to étale cohomology, $\ell$ invertible in $k$.</p> <p>This is translated into a statement of the comparison of the cohomology of motivic complexes with respect to different topologies.</p> <p>For $\ell = 2$, this is the Milnor conjecture. Even for this, Voevodsky got the Fields medal.</p> <p>For $n = 0$, this is obvious, for $n = 1$, it follows from Hilbert's Theorem 90. A corollary is that the Galois cohomology ring is generated by elements of degree $1$.</p> <p>It implies the Beilinson-Lichtenbaum conjecture, a more general conjecture for varieties, not just fields.</p> Answer 11: <p>In discrete math:</p> <p>Keevash's probabilisitic proof of Steiner's conjecture (On the Existence of Designs) is perhaps the biggest recent-ish result in combinatorics [due to the problem's ridiculously long pedigree].</p> <p>But if you allow a little further back, I think Szemeredi's regularity lemma would have to take the cake for its outrageous applicability.</p> Answer 12: <p>Probably one of the major developments in homotopy theory is the <a href="https://arxiv.org/abs/0908.3724">solution to the Kervaire invariant one problem</a> by Hill, Hopkins and Ravenel. It showcased the power of equivariant homotopy theory in solving a problem that a priori has nothing to do with it. In particular the idea to use the slice spectral sequence to get a handle on the homotopy fixed point spectral sequence of a certain spectrum is a real eye-opener.</p> Answer 13: <p>Perelman's proof of the Geometrization conjecture (see <a href="https://arxiv.org/abs/math/0211159" rel="nofollow noreferrer">here</a>, <a href="https://arxiv.org/abs/math/0303109" rel="nofollow noreferrer">here</a> and <a href="https://arxiv.org/abs/math/0307245" rel="nofollow noreferrer">here</a>) was the crowning achievement of decades of work. It was the most important of Thurston's conjectures about the topology of 3-manifolds.</p> <p>Oh, and I nearly forgot to mention that the Poincare Conjecture is a consequence.</p> Answer 14: <p>Agol's proof of the <a href="http://www.math.uiuc.edu/documenta/vol-18/vol-18-eng.html" rel="noreferrer">Virtual Haken conjecture</a> was a wonderful application of the tools developed by Wise and his coauthors in geometric group theory to 3-manifold topology. The Virtual Haken conjecture, which can be thought of as the topological classification of compact 3-manifolds, is a fundamental result.</p> <p>Agol's theorem:</p> <blockquote> <p>Every cubulable hyperbolic group is virtually special.</p> </blockquote> <p>also has other dramatic consequences for hyperbolic groups. It's arguably the most important theorem proved in both algebra <em>and</em> topology in the last five years.</p> Answer 15: <p>Kahn--Markovic's proofs of the <a href="https://arxiv.org/abs/0910.5501" rel="nofollow noreferrer">Surface Subgroup conjecture</a> and the <a href="https://arxiv.org/abs/1101.1330" rel="nofollow noreferrer">Ehrenpreis conjecture</a>.</p> Answer 16: <p>I'm going for:</p> <blockquote> <p><a href="http://mathworld.wolfram.com/GausssClassNumberProblem.html" rel="nofollow noreferrer">Gauss's Class Number Problem</a></p> </blockquote> <p>solved for $m\le100$ by Watkins in 2004, in '<a href="http://www.ams.org/journals/mcom/2004-73-246/S0025-5718-03-01517-5/S0025-5718-03-01517-5.pdf" rel="nofollow noreferrer">Class Numbers of Imaginary Quadratic Fields</a>'.</p> Answer 17: <p>Taniyama, Shimura and later Weil conjectured around 60 years ago that the L-function of an elliptic curve arises from a modular form. This conjecture was known to entail the last Fermat theorem after pioneering works of Hellegouarch, Frey and Ribet. In 1994, after a very long and technical work made in isolation and relying on Galois representations, Wiles managed to prove this conjecture in the semi-stable case, which was enough to establish the veracity of the now called Fermat-Wiles theorem. Based on this work, an article published in 1999 by Breuil, Conrad, Diamond and Taylor establish the result in full generality.</p> <p>This theorem is a tiny part of the Langlands program, dating from 1967.</p> Answer 18: <p>The proof of the <strong>Milnor-Bloch-Kato conjecture</strong> by <em>Vladimir Voevodsky</em> (Fields medal 2002):</p> <p>One has a norm residue isomorphism $$\mathrm{K}^\mathrm{M}_n(k)/\ell \to \mathrm{H}^n(k, \mu_\ell^{\otimes n})$$ from Milnor $K$-theory to étale cohomology, $\ell$ invertible in $k$.</p> <p>This is translated into a statement of the comparison of the cohomology of motivic complexes with respect to different topologies.</p> <p>For $\ell = 2$, this is the Milnor conjecture. Even for this, Voevodsky got the Fields medal.</p> <p>For $n = 0$, this is obvious, for $n = 1$, it follows from Hilbert's Theorem 90. A corollary is that the Galois cohomology ring is generated by elements of degree $1$.</p> <p>It implies the Beilinson-Lichtenbaum conjecture, a more general conjecture for varieties, not just fields.</p> Answer 19: <p>In discrete math:</p> <p>Keevash's probabilisitic proof of Steiner's conjecture (On the Existence of Designs) is perhaps the biggest recent-ish result in combinatorics [due to the problem's ridiculously long pedigree].</p> <p>But if you allow a little further back, I think Szemeredi's regularity lemma would have to take the cake for its outrageous applicability.</p> Answer 20: <p>Probably one of the major developments in homotopy theory is the <a href="https://arxiv.org/abs/0908.3724">solution to the Kervaire invariant one problem</a> by Hill, Hopkins and Ravenel. It showcased the power of equivariant homotopy theory in solving a problem that a priori has nothing to do with it. In particular the idea to use the slice spectral sequence to get a handle on the homotopy fixed point spectral sequence of a certain spectrum is a real eye-opener.</p> Answer 21: <p>Perelman's proof of the Geometrization conjecture (see <a href="https://arxiv.org/abs/math/0211159" rel="nofollow noreferrer">here</a>, <a href="https://arxiv.org/abs/math/0303109" rel="nofollow noreferrer">here</a> and <a href="https://arxiv.org/abs/math/0307245" rel="nofollow noreferrer">here</a>) was the crowning achievement of decades of work. It was the most important of Thurston's conjectures about the topology of 3-manifolds.</p> <p>Oh, and I nearly forgot to mention that the Poincare Conjecture is a consequence.</p> Answer 22: <p>Agol's proof of the <a href="http://www.math.uiuc.edu/documenta/vol-18/vol-18-eng.html" rel="noreferrer">Virtual Haken conjecture</a> was a wonderful application of the tools developed by Wise and his coauthors in geometric group theory to 3-manifold topology. The Virtual Haken conjecture, which can be thought of as the topological classification of compact 3-manifolds, is a fundamental result.</p> <p>Agol's theorem:</p> <blockquote> <p>Every cubulable hyperbolic group is virtually special.</p> </blockquote> <p>also has other dramatic consequences for hyperbolic groups. It's arguably the most important theorem proved in both algebra <em>and</em> topology in the last five years.</p> Answer 23: <p>Kahn--Markovic's proofs of the <a href="https://arxiv.org/abs/0910.5501" rel="nofollow noreferrer">Surface Subgroup conjecture</a> and the <a href="https://arxiv.org/abs/1101.1330" rel="nofollow noreferrer">Ehrenpreis conjecture</a>.</p> Answer 24: <p>I'm going for:</p> <blockquote> <p><a href="http://mathworld.wolfram.com/GausssClassNumberProblem.html" rel="nofollow noreferrer">Gauss's Class Number Problem</a></p> </blockquote> <p>solved for $m\le100$ by Watkins in 2004, in '<a href="http://www.ams.org/journals/mcom/2004-73-246/S0025-5718-03-01517-5/S0025-5718-03-01517-5.pdf" rel="nofollow noreferrer">Class Numbers of Imaginary Quadratic Fields</a>'.</p>
Looking for Arnol&#39;d quote about Russian students vs western mathematicians
259,909
10
https://mathoverflow.net/questions/259909/looking-for-arnold-quote-about-russian-students-vs-western-mathematicians
<p>I think I once saw a sentence in an article by V.I. Arnol'd saying something like: here is a problem that every Russian schoolchild can solve, but no western mathematician can solve. But I can't find it now. </p> <blockquote> <p>Does anyone know the quote and where it was written?</p> </blockquote> <p>I ask because I'm preparing a talk on research by high school students, and I might include this quote if the actual wording isn't too inflammatory.</p>
[ { "answer_id": 259911, "score": 15, "text": "<p>You possibly mean this 'two volumes' problem from the book ``Problems for children from 5 to 15'':</p>\n\n<p>Russian original:</p>\n\n<ol start=\"13\">\n<li>На книжной полке рядом стоят два тома Пушкина: первый и второй. Страницы каждого тома имеют вместе толщину 2 см, а обложка –– каждая –– 2 мм. Червь прогрыз (перпендикулярно страницам) от первой страницы первого тома до\nпоследней страницы второго тома. Какой путь он прогрыз?\n[Эта топологическая задача с невероятным ответом ––\n4 мм –– совершенно недоступна академикам, но некоторые\nдошкольники легко справляются с ней.]</li>\n</ol>\n\n<p>English translation:</p>\n\n<ol start=\"13\">\n<li>Two volumes of Pushkin, the first and the second, are side-by-side on\na bookshelf. The pages of each volume are 2 cm thick, and the cover – front\nand back each – is 2 mm. A bookworm has gnawed through (perpendicular to\nthe pages) from the first page of volume 1 to the last page of volume 2. How\nlong is the bookworm’s track?\n[This topological problem with an incredible answer – 4 mm – is absolutely\nimpossible for academicians, but some preschoolers handle it with ease.]</li>\n</ol>\n" }, { "answer_id": 259913, "score": 5, "text": "<p>Here's Arnold's essay on teaching mathematics.</p>\n\n<p><a href=\"http://pauli.uni-muenster.de/~munsteg/arnold.html\">http://pauli.uni-muenster.de/~munsteg/arnold.html</a></p>\n\n<p>\"For example, these students have never seen a paraboloid and a question on the form of the surface given by the equation $xy = z^2$ puts the mathematicians studying at ENS into a stupor. Drawing a curve given by parametric equations (like $x = t^3 - 3t$, $y = t^4 - 2t^2$) on a plane is a totally impossible problem for students (and, probably, even for most French professors of mathematics).\"</p>\n" }, { "answer_id": 259912, "score": 4, "text": "<p>Maybe you are referring to his <a href=\"https://imaginary.org/background-material/school-taskbook-from-5-to-15\" rel=\"nofollow noreferrer\">book</a> with problems for schoolchildren? See problem 13 there, and also a remark in the preface.</p>\n" }, { "answer_id": 259911, "score": 15, "text": "<p>You possibly mean this 'two volumes' problem from the book ``Problems for children from 5 to 15'':</p>\n\n<p>Russian original:</p>\n\n<ol start=\"13\">\n<li>На книжной полке рядом стоят два тома Пушкина: первый и второй. Страницы каждого тома имеют вместе толщину 2 см, а обложка –– каждая –– 2 мм. Червь прогрыз (перпендикулярно страницам) от первой страницы первого тома до\nпоследней страницы второго тома. Какой путь он прогрыз?\n[Эта топологическая задача с невероятным ответом ––\n4 мм –– совершенно недоступна академикам, но некоторые\nдошкольники легко справляются с ней.]</li>\n</ol>\n\n<p>English translation:</p>\n\n<ol start=\"13\">\n<li>Two volumes of Pushkin, the first and the second, are side-by-side on\na bookshelf. The pages of each volume are 2 cm thick, and the cover – front\nand back each – is 2 mm. A bookworm has gnawed through (perpendicular to\nthe pages) from the first page of volume 1 to the last page of volume 2. How\nlong is the bookworm’s track?\n[This topological problem with an incredible answer – 4 mm – is absolutely\nimpossible for academicians, but some preschoolers handle it with ease.]</li>\n</ol>\n" }, { "answer_id": 259913, "score": 5, "text": "<p>Here's Arnold's essay on teaching mathematics.</p>\n\n<p><a href=\"http://pauli.uni-muenster.de/~munsteg/arnold.html\">http://pauli.uni-muenster.de/~munsteg/arnold.html</a></p>\n\n<p>\"For example, these students have never seen a paraboloid and a question on the form of the surface given by the equation $xy = z^2$ puts the mathematicians studying at ENS into a stupor. Drawing a curve given by parametric equations (like $x = t^3 - 3t$, $y = t^4 - 2t^2$) on a plane is a totally impossible problem for students (and, probably, even for most French professors of mathematics).\"</p>\n" }, { "answer_id": 259912, "score": 4, "text": "<p>Maybe you are referring to his <a href=\"https://imaginary.org/background-material/school-taskbook-from-5-to-15\" rel=\"nofollow noreferrer\">book</a> with problems for schoolchildren? See problem 13 there, and also a remark in the preface.</p>\n" }, { "answer_id": 259911, "score": 15, "text": "<p>You possibly mean this 'two volumes' problem from the book ``Problems for children from 5 to 15'':</p>\n\n<p>Russian original:</p>\n\n<ol start=\"13\">\n<li>На книжной полке рядом стоят два тома Пушкина: первый и второй. Страницы каждого тома имеют вместе толщину 2 см, а обложка –– каждая –– 2 мм. Червь прогрыз (перпендикулярно страницам) от первой страницы первого тома до\nпоследней страницы второго тома. Какой путь он прогрыз?\n[Эта топологическая задача с невероятным ответом ––\n4 мм –– совершенно недоступна академикам, но некоторые\nдошкольники легко справляются с ней.]</li>\n</ol>\n\n<p>English translation:</p>\n\n<ol start=\"13\">\n<li>Two volumes of Pushkin, the first and the second, are side-by-side on\na bookshelf. The pages of each volume are 2 cm thick, and the cover – front\nand back each – is 2 mm. A bookworm has gnawed through (perpendicular to\nthe pages) from the first page of volume 1 to the last page of volume 2. How\nlong is the bookworm’s track?\n[This topological problem with an incredible answer – 4 mm – is absolutely\nimpossible for academicians, but some preschoolers handle it with ease.]</li>\n</ol>\n" }, { "answer_id": 259913, "score": 5, "text": "<p>Here's Arnold's essay on teaching mathematics.</p>\n\n<p><a href=\"http://pauli.uni-muenster.de/~munsteg/arnold.html\">http://pauli.uni-muenster.de/~munsteg/arnold.html</a></p>\n\n<p>\"For example, these students have never seen a paraboloid and a question on the form of the surface given by the equation $xy = z^2$ puts the mathematicians studying at ENS into a stupor. Drawing a curve given by parametric equations (like $x = t^3 - 3t$, $y = t^4 - 2t^2$) on a plane is a totally impossible problem for students (and, probably, even for most French professors of mathematics).\"</p>\n" }, { "answer_id": 259912, "score": 4, "text": "<p>Maybe you are referring to his <a href=\"https://imaginary.org/background-material/school-taskbook-from-5-to-15\" rel=\"nofollow noreferrer\">book</a> with problems for schoolchildren? See problem 13 there, and also a remark in the preface.</p>\n" } ]
Question: Looking for Arnol&#39;d quote about Russian students vs western mathematicians Answer 1: <p>You possibly mean this 'two volumes' problem from the book ``Problems for children from 5 to 15'':</p> <p>Russian original:</p> <ol start="13"> <li>На книжной полке рядом стоят два тома Пушкина: первый и второй. Страницы каждого тома имеют вместе толщину 2 см, а обложка –– каждая –– 2 мм. Червь прогрыз (перпендикулярно страницам) от первой страницы первого тома до последней страницы второго тома. Какой путь он прогрыз? [Эта топологическая задача с невероятным ответом –– 4 мм –– совершенно недоступна академикам, но некоторые дошкольники легко справляются с ней.]</li> </ol> <p>English translation:</p> <ol start="13"> <li>Two volumes of Pushkin, the first and the second, are side-by-side on a bookshelf. The pages of each volume are 2 cm thick, and the cover – front and back each – is 2 mm. A bookworm has gnawed through (perpendicular to the pages) from the first page of volume 1 to the last page of volume 2. How long is the bookworm’s track? [This topological problem with an incredible answer – 4 mm – is absolutely impossible for academicians, but some preschoolers handle it with ease.]</li> </ol> Answer 2: <p>Here's Arnold's essay on teaching mathematics.</p> <p><a href="http://pauli.uni-muenster.de/~munsteg/arnold.html">http://pauli.uni-muenster.de/~munsteg/arnold.html</a></p> <p>"For example, these students have never seen a paraboloid and a question on the form of the surface given by the equation $xy = z^2$ puts the mathematicians studying at ENS into a stupor. Drawing a curve given by parametric equations (like $x = t^3 - 3t$, $y = t^4 - 2t^2$) on a plane is a totally impossible problem for students (and, probably, even for most French professors of mathematics)."</p> Answer 3: <p>Maybe you are referring to his <a href="https://imaginary.org/background-material/school-taskbook-from-5-to-15" rel="nofollow noreferrer">book</a> with problems for schoolchildren? See problem 13 there, and also a remark in the preface.</p> Answer 4: <p>You possibly mean this 'two volumes' problem from the book ``Problems for children from 5 to 15'':</p> <p>Russian original:</p> <ol start="13"> <li>На книжной полке рядом стоят два тома Пушкина: первый и второй. Страницы каждого тома имеют вместе толщину 2 см, а обложка –– каждая –– 2 мм. Червь прогрыз (перпендикулярно страницам) от первой страницы первого тома до последней страницы второго тома. Какой путь он прогрыз? [Эта топологическая задача с невероятным ответом –– 4 мм –– совершенно недоступна академикам, но некоторые дошкольники легко справляются с ней.]</li> </ol> <p>English translation:</p> <ol start="13"> <li>Two volumes of Pushkin, the first and the second, are side-by-side on a bookshelf. The pages of each volume are 2 cm thick, and the cover – front and back each – is 2 mm. A bookworm has gnawed through (perpendicular to the pages) from the first page of volume 1 to the last page of volume 2. How long is the bookworm’s track? [This topological problem with an incredible answer – 4 mm – is absolutely impossible for academicians, but some preschoolers handle it with ease.]</li> </ol> Answer 5: <p>Here's Arnold's essay on teaching mathematics.</p> <p><a href="http://pauli.uni-muenster.de/~munsteg/arnold.html">http://pauli.uni-muenster.de/~munsteg/arnold.html</a></p> <p>"For example, these students have never seen a paraboloid and a question on the form of the surface given by the equation $xy = z^2$ puts the mathematicians studying at ENS into a stupor. Drawing a curve given by parametric equations (like $x = t^3 - 3t$, $y = t^4 - 2t^2$) on a plane is a totally impossible problem for students (and, probably, even for most French professors of mathematics)."</p> Answer 6: <p>Maybe you are referring to his <a href="https://imaginary.org/background-material/school-taskbook-from-5-to-15" rel="nofollow noreferrer">book</a> with problems for schoolchildren? See problem 13 there, and also a remark in the preface.</p> Answer 7: <p>You possibly mean this 'two volumes' problem from the book ``Problems for children from 5 to 15'':</p> <p>Russian original:</p> <ol start="13"> <li>На книжной полке рядом стоят два тома Пушкина: первый и второй. Страницы каждого тома имеют вместе толщину 2 см, а обложка –– каждая –– 2 мм. Червь прогрыз (перпендикулярно страницам) от первой страницы первого тома до последней страницы второго тома. Какой путь он прогрыз? [Эта топологическая задача с невероятным ответом –– 4 мм –– совершенно недоступна академикам, но некоторые дошкольники легко справляются с ней.]</li> </ol> <p>English translation:</p> <ol start="13"> <li>Two volumes of Pushkin, the first and the second, are side-by-side on a bookshelf. The pages of each volume are 2 cm thick, and the cover – front and back each – is 2 mm. A bookworm has gnawed through (perpendicular to the pages) from the first page of volume 1 to the last page of volume 2. How long is the bookworm’s track? [This topological problem with an incredible answer – 4 mm – is absolutely impossible for academicians, but some preschoolers handle it with ease.]</li> </ol> Answer 8: <p>Here's Arnold's essay on teaching mathematics.</p> <p><a href="http://pauli.uni-muenster.de/~munsteg/arnold.html">http://pauli.uni-muenster.de/~munsteg/arnold.html</a></p> <p>"For example, these students have never seen a paraboloid and a question on the form of the surface given by the equation $xy = z^2$ puts the mathematicians studying at ENS into a stupor. Drawing a curve given by parametric equations (like $x = t^3 - 3t$, $y = t^4 - 2t^2$) on a plane is a totally impossible problem for students (and, probably, even for most French professors of mathematics)."</p> Answer 9: <p>Maybe you are referring to his <a href="https://imaginary.org/background-material/school-taskbook-from-5-to-15" rel="nofollow noreferrer">book</a> with problems for schoolchildren? See problem 13 there, and also a remark in the preface.</p>
Jacobi&#39;s elliptic functions and plane sections of a torus
260,202
10
https://mathoverflow.net/questions/260202/jacobis-elliptic-functions-and-plane-sections-of-a-torus
<p>In <span class="math-container">$\mathbb R^3$</span> with Cartesian coordinates <span class="math-container">$(x,y,z),$</span> revolve the circle <span class="math-container">$(x-\sqrt 2)^2+z^2 =1,\ y=0$</span> about the <span class="math-container">$z$</span>-axis.</p> <p>This yields a torus embedded in <span class="math-container">$3$</span>-space that is conformally equivalent to the flat torus <span class="math-container">$(\mathbb R/(2\pi))^2,$</span> via a mapping that takes <span class="math-container">$(x,y,z)=(1+\sqrt 2,0,0)$</span> to <span class="math-container">$(0,0)$</span> and <span class="math-container">$(x,z) = (-1-\sqrt 2,0, 0)$</span> to <span class="math-container">$(0,\pi),$</span> and <span class="math-container">$(-1+\sqrt 2,0,0)$</span> to <span class="math-container">$(\pi,0)$</span> and <span class="math-container">$(1-\sqrt 2,0,0)$</span> to <span class="math-container">$(\pi,\pi).$</span></p> <p>One of Jacobi's elliptic functions gives us a <span class="math-container">$2$</span>-to-<span class="math-container">$1$</span> mapping from the flat torus into <span class="math-container">$\mathbb C\cup\{\infty\}$</span> that takes <span class="math-container">$(0,0)$</span> to <span class="math-container">$0$</span> and <span class="math-container">$(0,\pi)$</span> to <span class="math-container">$\infty.$</span></p> <p>Then the inverse of the stereographic projection takes <span class="math-container">$\mathbb C\cup\{\infty\}$</span> to the sphere <span class="math-container">$x^2+y^2+z^2=1$</span> while taking <span class="math-container">$0$</span> to <span class="math-container">$(1,0,0)$</span> and <span class="math-container">$\infty$</span> to <span class="math-container">$(-1,0,0).$</span></p> <p>The composition of mappings takes the embedded torus to the sphere.</p> <p><b>Fact 1:</b> This composition of mappings takes the largest and the smallest of the four <span class="math-container">$x$</span>-intercepts of the torus respectively to the largest and smallest of the two <span class="math-container">$x$</span>-intercepts of the sphere.</p> <p><b>Fact 2:</b> This composition of mappings is conformal except at four points on the torus (corresponding to points at which the derivative of Jacobi's function is <span class="math-container">$0$</span>).</p> <p>(Unless I got some of the nitpicking details wrong about which points map to which) I believe I have reason to suspect the following really startling proposition:</p> <blockquote> <p>The intersection of this sphere with each plane containing the <span class="math-container">$x$</span>-axis is the image under this composition of mappings of the intersection of the torus with the <b>same</b> plane.</p> </blockquote> <p>So I'm wondering:</p> <ul> <li>Is this true?</li> <li>Is it known?</li> <li>Can it be proved by using some known fact about Jacobi's function and maybe some elbow grease?</li> </ul> <p><b>APPENDIX:</b> The inverse of the conformal mapping from the embedded torus to the flat torus is this: <span class="math-container">\begin{align} \text{For }\alpha,\beta\in\mathbb R/(2\pi), \\[4pt] x &amp; = \cos\beta/(\sqrt2-\cos\alpha) \\[4pt] y &amp; = \sin\beta/(\sqrt2-\cos\alpha) \\[4pt] z &amp; = \sin\alpha/(\sqrt2-\cos\alpha) \\[4pt] &amp; {} \qquad \uparrow \\ &amp; {} \qquad \text{same denominator in all three coordinates} \end{align}</span></p> <p><b>POSTSCRIPT:</b> If I'm not mistaken, I've reduced this to the following.</p> <p>You have a doubly periodic holomorphic function with periods <span class="math-container">$2\pi$</span> and <span class="math-container">$2\pi i$</span> that takes <span class="math-container">$0$</span> and <span class="math-container">$\pi + i\pi$</span> to <span class="math-container">$0$</span> and takes <span class="math-container">$\pi$</span> and <span class="math-container">$i\pi$</span> to <span class="math-container">$\infty.$</span></p> <p>Now look at the inverse-images of straight lines through <span class="math-container">$0$</span> corresponding to arguments (i.e. angles from the real axis) between <span class="math-container">$0$</span> and <span class="math-container">$45^\circ$</span>. Those inverse-images should be graphs of functions of the form <span class="math-container">$y=\arcsin(c\sin(x))$</span> where <span class="math-container">$x$</span> and <span class="math-container">$y$</span> are respectively the real and imaginary parts.</p> <p>Is that a known proposition about Jacobi's functions?</p> <p><b>Postscript:</b> Greg Egan created a visual illustration of what I described above. What is seen below is a compressed version. His original looks better. But it still moves as fast as this one. A version that moves much more slowly would be better.</p> <p><a href="https://i.stack.imgur.com/0wX0I.gif" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/0wX0I.gif" alt="Torus globe" /></a></p>
[ { "answer_id": 261263, "score": 12, "text": "<p>This is more comment and question than answer, but I'm writing it as an answer because it's too long to fit as a comment.</p>\n\n<p>Michael Hardy kindly pointed me to a reference on the original mapping between the plane and the sphere:</p>\n\n<p><a href=\"http://www.jstor.org/stable/2369491\" rel=\"noreferrer\">“A Quincuncial Projection of the Sphere”</a>, C. S. Peirce, <em>American Journal of Mathematics</em>, Vol. 2, No. 4 (Dec., 1879), pp. 394-396</p>\n\n<p>Here, Peirce defines the mapping between a point $(x,y)$ in the plane and a point on the sphere with co-latitude $p$ and longitude $\\theta$ as:</p>\n\n<p>$$\\operatorname{cn}\\left(x + i y \\mid \\frac{1}{2}\\right) = \\tan\\left(\\frac{p}{2} \\right)\\left( \\cos(\\theta) + i \\sin(\\theta) \\right)$$</p>\n\n<p>This $\\operatorname{cn}$ function has periods $4 K$ and $2(1+i)K$, where $K=F\\left(\\frac{1}{2}\\right)$. It has zeroes at $(2q+1) K + 2r K i$ and poles at $2q K + (2r+1) K i$, for integers $q, r$. To be clear, the parameters for the Jacobi $\\text{cn}$ function and the complete elliptic integral $F$ are what is usually called $m$, as opposed to the modulus, usually called $k$, where $m=k^2$.</p>\n\n<p>The tables in Peirce are rescaled to make the factor of $K$ go away, but they also seem to be flipped and translated relative to his initial definition. I can reproduce Peirce’s tables with Mathematica if I take them to be describing $x',y'$ where, in relation to his original $x, y$ we have:</p>\n\n<p>$$x'+ i y' = 1 - \\frac{(x + i y)}{K}$$</p>\n\n<p>Now, Michael Hardy requires a function on $(\\mathbb{R}/(2\\pi))^2$, construed as a square in the complex plane, that has periods $2\\pi$ and $2 \\pi i$, zeroes at $0$ and $\\pi+i \\pi$ and poles at $\\pi$ and $i \\pi$. We can write this as:</p>\n\n<p>$$H(z) = \\operatorname{cn}\\left(K \\left(\\left(\\frac{1+i}{\\pi}\\right) z + 1 \\right) \\mid \\frac{1}{2}\\right)$$</p>\n\n<p>He defines a conformal map between a torus in $\\mathbb{R}^3$ and this square in the complex plane via:</p>\n\n<p>$$\\phi_0^{-1}(\\alpha + \\beta i) = \\frac{(\\cos(\\beta),\\sin(\\beta),\\sin(\\alpha))}{\\sqrt{2}-\\cos(\\alpha)}$$</p>\n\n<p>We also have the stereographic map, projecting from the south pole of the unit sphere onto a plane through the equator, whose inverse can be written:</p>\n\n<p>$$\\psi_0^{-1}\\left(\\tan\\left(\\frac{p}{2}\\right)\\left( \\cos(\\theta) + i \\sin(\\theta) \\right)\\right) = (\\cos(\\theta) \\sin(p), \\sin(\\theta) \\sin(p), \\cos(p))$$</p>\n\n<p>[<strong>Edited to add, 21 Feb 2017:</strong> I’ve just belatedly noticed that in the question, the stereographic projection is defined with the poles of the unit sphere on the $x$-axis, rather than the usual convention of putting the poles on the $z$-axis. This explains why I was unable to verify the original claims, and ended up reorienting the torus to make things work.]</p>\n\n<p>Now, the conjecture is that there is a family of planes $\\scr{P}$ such that for any plane $P \\in \\scr{P}$ we have:</p>\n\n<p>$$\\psi^{-1}(H(\\phi(\\scr{T} \\cap P))) = \\scr{S} \\cap P$$</p>\n\n<p>where $\\scr{T}$ is the torus, $\\scr{S}$ is the sphere, and I've dropped the subscripts $0$ from the maps because I want to allow the original versions to be adjusted slightly.</p>\n\n<p>To see how to adjust these maps, note that we want $\\scr{P}$ to include both the $x y$-plane and the $x z$-plane. The stereographic projection of the equator of the sphere is just the unit circle in the complex plane, and the preimage of the unit circle under $H$ consists of the four lines $\\alpha=\\frac{\\pi}{2}$, $\\alpha=\\frac{3\\pi}{2}$, $\\beta=\\frac{\\pi}{2}$ and $\\beta=\\frac{3\\pi}{2}$, where we’re writing points in the fundamental domain as $z=\\alpha + \\beta i$.</p>\n\n<p>The map $\\phi_0^{-1}$ from the fundamental domain to the torus maps the lines $\\alpha=0$ and $\\alpha=\\pi$ to the two circles on the torus that lie in the $x y$-plane. So we need to define:</p>\n\n<p>$$\\phi^{-1}(\\alpha + \\beta i) = \\phi_0^{-1}\\left(\\left(\\alpha + \\frac{\\pi}{2}\\right) + \\beta i\\right) = \\frac{(\\cos(\\beta),\\sin(\\beta),\\cos(\\alpha))}{\\sqrt{2}+\\sin(\\alpha)}$$</p>\n\n<p>To accommodate the $x z$-plane, note that the intersection of the torus with this plane corresponds to the lines $\\beta=0$ and $\\beta=\\pi$, the first of which is mapped by the initial linear transformation in $H$ to a line at 45 degrees that passes through the point $z=K$ in the domain of the original Jacobi function, and contains alternating zeroes and poles of the function. If we take a midpoint between a zero and a pole, such as $\\frac{3}{2}K+\\frac{1}{2}K i$, the Jacobi function evaluates to $-\\sqrt{\\frac{1}{2}}(1+i)$, which corresponds to a longitude of $\\frac{5\\pi}{4}$ on the sphere. We can’t remedy this by offsetting $\\beta$ in the torus map, because the <em>only</em> lines of constant $\\beta$ in the fundamental domain of $H$ that ultimately map to meridians on the sphere are $\\beta=0$ and $\\beta=\\pi$. So we need to redefine the map from the complex plane to the sphere by adding a rotation of $-\\frac{\\pi}{4}$:</p>\n\n<p>$$\\psi^{-1}(z) = \\psi_0^{-1}(\\exp(-i \\pi/4) z)$$</p>\n\n<p>Having made these changes, let’s see exactly what happens for the two planes.</p>\n\n<p>The lines $\\alpha=\\frac{\\pi}{2}$ and $\\alpha=\\frac{3\\pi}{2}$ clearly map under $\\phi^{-1}$ to the two circles in the $x y$-plane where it intersects the torus $\\scr{T}$. Under $H$, they map 2-to-1 to two opposite quadrants of the unit circle (those with arguments between $-\\frac{\\pi}{2}$ and $-\\pi$, and between $0$ and $\\frac{\\pi}{2}$), and then under $\\psi^{-1}$ to two opposite quadrants of the equator. So they don't map to the <em>complete</em> equator; the preimage of the full equator in the fundamental domain of $H$ includes $\\beta=\\frac{\\pi}{2}$ and $\\beta=\\frac{3\\pi}{2}$. But certainly they map into the same plane as the image on the torus.</p>\n\n<p>For the lines $\\beta=0$ and $\\beta=\\pi$, it’s easy to see that for the torus $\\phi^{-1}$ gives us two circles in the $x z$-plane. Under $H$ and $\\psi^{-1}$, the two lines each separately cover the entire circle where the sphere intersects the $x z$-plane. So in this case, the conjecture is exactly true.</p>\n\n<p>However, given the maps that seem to have been forced on us by the need to make things work for these two planes, I can’t see how to make the case $x=\\pm z$ work, or $y =\\pm z$ (I'm not sure whether the claim that the conjecture is true for the former is a typo and the latter was intended, because the initial conjecture was for planes containing the $x$-axis, and $x=\\pm z$ do not contain the $x$-axis).</p>\n\n<p>For the case $y =\\pm z$, consider the curve (in fact, a straight line) in the fundamental domain of $H$ given by:</p>\n\n<p>$$\\alpha = \\cos^{-1}(\\gamma)$$\n$$\\beta = \\sin^{-1}(\\gamma)$$</p>\n\n<p>Since $\\sin(\\beta)$ and $\\cos(\\alpha)$ will be equal, the image of this line under $\\phi^{-1}$ lies in the plane $y=z$.</p>\n\n<p>Suppose we put $\\gamma=\\frac{1}{\\sqrt{2}}$, giving us the point:</p>\n\n<p>$$p_0 = \\frac{(1+i)\\pi}{4}$$</p>\n\n<p>We have:</p>\n\n<p>$$\\phi^{-1}(p_0) = \\left(\\frac{1}{3}, \\frac{1}{3}, \\frac{1}{3}\\right)$$</p>\n\n<p>However:</p>\n\n<p>$$H(p_0) = -\\sqrt{\\sqrt{2}-1} i$$\n$$\\psi^{-1}(H(p_0)) = \\left(-\\sqrt{\\sqrt{2}-1},-\\sqrt{\\sqrt{2}-1},\\sqrt{2}-1\\right)$$</p>\n\n<p>which is not in the plane $y=z$.</p>\n\n<p><strong>Edited to add:</strong></p>\n\n<p>These plots might be helpful. They show the $(\\alpha, \\beta)$ domain, coloured according to which plane the image of each point lies on, when sent either to the sphere or to the torus. Here, I am using the <strong>original</strong> maps, $\\psi_0^{-1} H$ and $\\phi_0^{-1}$ to take the $(\\alpha, \\beta)$ domain to the surfaces of the sphere and the torus. The hue of each point depends on the angle measured around the specified axis in $\\mathbb{R}^3$, modulo $\\pi$, with points coloured black when the angle lies close to a multiple of $\\pi/4$.</p>\n\n<p><a href=\"https://i.stack.imgur.com/wUESs.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/wUESs.png\" alt=\"Planes through (1,0,0)\"></a>\n<a href=\"https://i.stack.imgur.com/1JgIG.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/1JgIG.png\" alt=\"Planes through (0,1,0)\"></a>\n<a href=\"https://i.stack.imgur.com/7UquT.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/7UquT.png\" alt=\"Planes through (0,0,1)\"></a>\n<a href=\"https://i.stack.imgur.com/og2bC.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/og2bC.png\" alt=\"Planes through (1,1,0)\"></a>\n<a href=\"https://i.stack.imgur.com/rRqNh.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/rRqNh.png\" alt=\"Planes through (1,0,1)\"></a>\n<a href=\"https://i.stack.imgur.com/LaHKE.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/LaHKE.png\" alt=\"Planes through (0,1,1)\"></a></p>\n\n<p><strong>Edited to add:</strong> Noting the similarity between the plot for the torus with planes around the $x$-axis and the plot for the sphere with planes around the $z$-axis, I've found that the two can be brought into very close alignment by a change of basis for the torus that preserves the axis, but measures the angles in the opposite direction and translates them all by $\\pi/4$. To be precise, if we use the standard basis for $\\mathbb{R}^3$ for the sphere, and the basis:</p>\n\n<p>$$\\left\\{\\left(0,\\frac{1}{\\sqrt{2}},-\\frac{1}{\\sqrt{2}}\\right),\\left(0,-\\frac{1}{\\sqrt{2}},-\\frac{1}{\\sqrt{2}}\\right),\\left(-1,0,0\\right)\\right\\}$$</p>\n\n<p>for the torus, and then compare the azimuthal angles in polar coordinates adapted to the respective bases, then these angles are very similar, differing from each other by an average of about 0.025.</p>\n\n<p>And the angles agree <strong>exactly</strong> when they are multiples of $\\pi/4$.</p>\n\n<p>I guess this must correspond to Michael Hardy's original observation.</p>\n\n<p><strong>Edited to add:</strong> This is what the final map from the torus to the sphere looks like:</p>\n\n<p><a href=\"https://i.stack.imgur.com/rE5UT.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/rE5UT.png\" alt=\"Map from torus to sphere\"></a></p>\n\n<p>Meridians pulled back from the sphere are in black, lines of latitude in red. The pullbacks of the meridians come very close to lying in the same planes as the original meridians on the sphere, if we define:</p>\n\n<p>$$\\phi_1^{-1}(\\alpha + \\beta i) = \\frac{(\\frac{1}{\\sqrt{2}}(\\sin(\\alpha)-\\sin(\\beta)),\\frac{1}{\\sqrt{2}}(\\sin(\\alpha)+\\sin(\\beta)),\\cos(\\beta))}{\\cos(\\alpha)-\\sqrt{2}}$$</p>\n\n<p>The 2-to-1 map from the torus to the sphere is then $\\psi_0^{-1} H \\phi_1$, where $\\psi_0^{-1}$ and $H$ are as previously defined.</p>\n" }, { "answer_id": 260781, "score": 2, "text": "<p>I hazarded this guess by</p>\n\n<ul>\n<li>looking at some crude pictures and at some published graphs that were almost certainly quite accurate, and</li>\n<li>seeing that the proposition is true in the two trivial cases, which are the $xy$-plane and the $xz$-plane, and</li>\n<li>seeing that it is also true of the planes $x=\\pm z$; this follows from a theorem of geometry published in the 1890s that is stronger than Villarceau's theorem. That stronger theorem says that certain loxodromes of any standard embedded torus lie in a plane.</li>\n</ul>\n\n<p>However yesterday I went over this the old-fashioned way, using some tables that were published in the 19th century. And a calculator for values of the arcsine. The proposition as stated is false, but comes fairly close to the truth numerically. Impressive counterfeits, but detectable if you're careful. Now I'm wondering if a slight modification of the curve that gets revolved about the $z$-axis would make this true.</p>\n" }, { "answer_id": 261263, "score": 12, "text": "<p>This is more comment and question than answer, but I'm writing it as an answer because it's too long to fit as a comment.</p>\n\n<p>Michael Hardy kindly pointed me to a reference on the original mapping between the plane and the sphere:</p>\n\n<p><a href=\"http://www.jstor.org/stable/2369491\" rel=\"noreferrer\">“A Quincuncial Projection of the Sphere”</a>, C. S. Peirce, <em>American Journal of Mathematics</em>, Vol. 2, No. 4 (Dec., 1879), pp. 394-396</p>\n\n<p>Here, Peirce defines the mapping between a point $(x,y)$ in the plane and a point on the sphere with co-latitude $p$ and longitude $\\theta$ as:</p>\n\n<p>$$\\operatorname{cn}\\left(x + i y \\mid \\frac{1}{2}\\right) = \\tan\\left(\\frac{p}{2} \\right)\\left( \\cos(\\theta) + i \\sin(\\theta) \\right)$$</p>\n\n<p>This $\\operatorname{cn}$ function has periods $4 K$ and $2(1+i)K$, where $K=F\\left(\\frac{1}{2}\\right)$. It has zeroes at $(2q+1) K + 2r K i$ and poles at $2q K + (2r+1) K i$, for integers $q, r$. To be clear, the parameters for the Jacobi $\\text{cn}$ function and the complete elliptic integral $F$ are what is usually called $m$, as opposed to the modulus, usually called $k$, where $m=k^2$.</p>\n\n<p>The tables in Peirce are rescaled to make the factor of $K$ go away, but they also seem to be flipped and translated relative to his initial definition. I can reproduce Peirce’s tables with Mathematica if I take them to be describing $x',y'$ where, in relation to his original $x, y$ we have:</p>\n\n<p>$$x'+ i y' = 1 - \\frac{(x + i y)}{K}$$</p>\n\n<p>Now, Michael Hardy requires a function on $(\\mathbb{R}/(2\\pi))^2$, construed as a square in the complex plane, that has periods $2\\pi$ and $2 \\pi i$, zeroes at $0$ and $\\pi+i \\pi$ and poles at $\\pi$ and $i \\pi$. We can write this as:</p>\n\n<p>$$H(z) = \\operatorname{cn}\\left(K \\left(\\left(\\frac{1+i}{\\pi}\\right) z + 1 \\right) \\mid \\frac{1}{2}\\right)$$</p>\n\n<p>He defines a conformal map between a torus in $\\mathbb{R}^3$ and this square in the complex plane via:</p>\n\n<p>$$\\phi_0^{-1}(\\alpha + \\beta i) = \\frac{(\\cos(\\beta),\\sin(\\beta),\\sin(\\alpha))}{\\sqrt{2}-\\cos(\\alpha)}$$</p>\n\n<p>We also have the stereographic map, projecting from the south pole of the unit sphere onto a plane through the equator, whose inverse can be written:</p>\n\n<p>$$\\psi_0^{-1}\\left(\\tan\\left(\\frac{p}{2}\\right)\\left( \\cos(\\theta) + i \\sin(\\theta) \\right)\\right) = (\\cos(\\theta) \\sin(p), \\sin(\\theta) \\sin(p), \\cos(p))$$</p>\n\n<p>[<strong>Edited to add, 21 Feb 2017:</strong> I’ve just belatedly noticed that in the question, the stereographic projection is defined with the poles of the unit sphere on the $x$-axis, rather than the usual convention of putting the poles on the $z$-axis. This explains why I was unable to verify the original claims, and ended up reorienting the torus to make things work.]</p>\n\n<p>Now, the conjecture is that there is a family of planes $\\scr{P}$ such that for any plane $P \\in \\scr{P}$ we have:</p>\n\n<p>$$\\psi^{-1}(H(\\phi(\\scr{T} \\cap P))) = \\scr{S} \\cap P$$</p>\n\n<p>where $\\scr{T}$ is the torus, $\\scr{S}$ is the sphere, and I've dropped the subscripts $0$ from the maps because I want to allow the original versions to be adjusted slightly.</p>\n\n<p>To see how to adjust these maps, note that we want $\\scr{P}$ to include both the $x y$-plane and the $x z$-plane. The stereographic projection of the equator of the sphere is just the unit circle in the complex plane, and the preimage of the unit circle under $H$ consists of the four lines $\\alpha=\\frac{\\pi}{2}$, $\\alpha=\\frac{3\\pi}{2}$, $\\beta=\\frac{\\pi}{2}$ and $\\beta=\\frac{3\\pi}{2}$, where we’re writing points in the fundamental domain as $z=\\alpha + \\beta i$.</p>\n\n<p>The map $\\phi_0^{-1}$ from the fundamental domain to the torus maps the lines $\\alpha=0$ and $\\alpha=\\pi$ to the two circles on the torus that lie in the $x y$-plane. So we need to define:</p>\n\n<p>$$\\phi^{-1}(\\alpha + \\beta i) = \\phi_0^{-1}\\left(\\left(\\alpha + \\frac{\\pi}{2}\\right) + \\beta i\\right) = \\frac{(\\cos(\\beta),\\sin(\\beta),\\cos(\\alpha))}{\\sqrt{2}+\\sin(\\alpha)}$$</p>\n\n<p>To accommodate the $x z$-plane, note that the intersection of the torus with this plane corresponds to the lines $\\beta=0$ and $\\beta=\\pi$, the first of which is mapped by the initial linear transformation in $H$ to a line at 45 degrees that passes through the point $z=K$ in the domain of the original Jacobi function, and contains alternating zeroes and poles of the function. If we take a midpoint between a zero and a pole, such as $\\frac{3}{2}K+\\frac{1}{2}K i$, the Jacobi function evaluates to $-\\sqrt{\\frac{1}{2}}(1+i)$, which corresponds to a longitude of $\\frac{5\\pi}{4}$ on the sphere. We can’t remedy this by offsetting $\\beta$ in the torus map, because the <em>only</em> lines of constant $\\beta$ in the fundamental domain of $H$ that ultimately map to meridians on the sphere are $\\beta=0$ and $\\beta=\\pi$. So we need to redefine the map from the complex plane to the sphere by adding a rotation of $-\\frac{\\pi}{4}$:</p>\n\n<p>$$\\psi^{-1}(z) = \\psi_0^{-1}(\\exp(-i \\pi/4) z)$$</p>\n\n<p>Having made these changes, let’s see exactly what happens for the two planes.</p>\n\n<p>The lines $\\alpha=\\frac{\\pi}{2}$ and $\\alpha=\\frac{3\\pi}{2}$ clearly map under $\\phi^{-1}$ to the two circles in the $x y$-plane where it intersects the torus $\\scr{T}$. Under $H$, they map 2-to-1 to two opposite quadrants of the unit circle (those with arguments between $-\\frac{\\pi}{2}$ and $-\\pi$, and between $0$ and $\\frac{\\pi}{2}$), and then under $\\psi^{-1}$ to two opposite quadrants of the equator. So they don't map to the <em>complete</em> equator; the preimage of the full equator in the fundamental domain of $H$ includes $\\beta=\\frac{\\pi}{2}$ and $\\beta=\\frac{3\\pi}{2}$. But certainly they map into the same plane as the image on the torus.</p>\n\n<p>For the lines $\\beta=0$ and $\\beta=\\pi$, it’s easy to see that for the torus $\\phi^{-1}$ gives us two circles in the $x z$-plane. Under $H$ and $\\psi^{-1}$, the two lines each separately cover the entire circle where the sphere intersects the $x z$-plane. So in this case, the conjecture is exactly true.</p>\n\n<p>However, given the maps that seem to have been forced on us by the need to make things work for these two planes, I can’t see how to make the case $x=\\pm z$ work, or $y =\\pm z$ (I'm not sure whether the claim that the conjecture is true for the former is a typo and the latter was intended, because the initial conjecture was for planes containing the $x$-axis, and $x=\\pm z$ do not contain the $x$-axis).</p>\n\n<p>For the case $y =\\pm z$, consider the curve (in fact, a straight line) in the fundamental domain of $H$ given by:</p>\n\n<p>$$\\alpha = \\cos^{-1}(\\gamma)$$\n$$\\beta = \\sin^{-1}(\\gamma)$$</p>\n\n<p>Since $\\sin(\\beta)$ and $\\cos(\\alpha)$ will be equal, the image of this line under $\\phi^{-1}$ lies in the plane $y=z$.</p>\n\n<p>Suppose we put $\\gamma=\\frac{1}{\\sqrt{2}}$, giving us the point:</p>\n\n<p>$$p_0 = \\frac{(1+i)\\pi}{4}$$</p>\n\n<p>We have:</p>\n\n<p>$$\\phi^{-1}(p_0) = \\left(\\frac{1}{3}, \\frac{1}{3}, \\frac{1}{3}\\right)$$</p>\n\n<p>However:</p>\n\n<p>$$H(p_0) = -\\sqrt{\\sqrt{2}-1} i$$\n$$\\psi^{-1}(H(p_0)) = \\left(-\\sqrt{\\sqrt{2}-1},-\\sqrt{\\sqrt{2}-1},\\sqrt{2}-1\\right)$$</p>\n\n<p>which is not in the plane $y=z$.</p>\n\n<p><strong>Edited to add:</strong></p>\n\n<p>These plots might be helpful. They show the $(\\alpha, \\beta)$ domain, coloured according to which plane the image of each point lies on, when sent either to the sphere or to the torus. Here, I am using the <strong>original</strong> maps, $\\psi_0^{-1} H$ and $\\phi_0^{-1}$ to take the $(\\alpha, \\beta)$ domain to the surfaces of the sphere and the torus. The hue of each point depends on the angle measured around the specified axis in $\\mathbb{R}^3$, modulo $\\pi$, with points coloured black when the angle lies close to a multiple of $\\pi/4$.</p>\n\n<p><a href=\"https://i.stack.imgur.com/wUESs.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/wUESs.png\" alt=\"Planes through (1,0,0)\"></a>\n<a href=\"https://i.stack.imgur.com/1JgIG.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/1JgIG.png\" alt=\"Planes through (0,1,0)\"></a>\n<a href=\"https://i.stack.imgur.com/7UquT.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/7UquT.png\" alt=\"Planes through (0,0,1)\"></a>\n<a href=\"https://i.stack.imgur.com/og2bC.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/og2bC.png\" alt=\"Planes through (1,1,0)\"></a>\n<a href=\"https://i.stack.imgur.com/rRqNh.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/rRqNh.png\" alt=\"Planes through (1,0,1)\"></a>\n<a href=\"https://i.stack.imgur.com/LaHKE.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/LaHKE.png\" alt=\"Planes through (0,1,1)\"></a></p>\n\n<p><strong>Edited to add:</strong> Noting the similarity between the plot for the torus with planes around the $x$-axis and the plot for the sphere with planes around the $z$-axis, I've found that the two can be brought into very close alignment by a change of basis for the torus that preserves the axis, but measures the angles in the opposite direction and translates them all by $\\pi/4$. To be precise, if we use the standard basis for $\\mathbb{R}^3$ for the sphere, and the basis:</p>\n\n<p>$$\\left\\{\\left(0,\\frac{1}{\\sqrt{2}},-\\frac{1}{\\sqrt{2}}\\right),\\left(0,-\\frac{1}{\\sqrt{2}},-\\frac{1}{\\sqrt{2}}\\right),\\left(-1,0,0\\right)\\right\\}$$</p>\n\n<p>for the torus, and then compare the azimuthal angles in polar coordinates adapted to the respective bases, then these angles are very similar, differing from each other by an average of about 0.025.</p>\n\n<p>And the angles agree <strong>exactly</strong> when they are multiples of $\\pi/4$.</p>\n\n<p>I guess this must correspond to Michael Hardy's original observation.</p>\n\n<p><strong>Edited to add:</strong> This is what the final map from the torus to the sphere looks like:</p>\n\n<p><a href=\"https://i.stack.imgur.com/rE5UT.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/rE5UT.png\" alt=\"Map from torus to sphere\"></a></p>\n\n<p>Meridians pulled back from the sphere are in black, lines of latitude in red. The pullbacks of the meridians come very close to lying in the same planes as the original meridians on the sphere, if we define:</p>\n\n<p>$$\\phi_1^{-1}(\\alpha + \\beta i) = \\frac{(\\frac{1}{\\sqrt{2}}(\\sin(\\alpha)-\\sin(\\beta)),\\frac{1}{\\sqrt{2}}(\\sin(\\alpha)+\\sin(\\beta)),\\cos(\\beta))}{\\cos(\\alpha)-\\sqrt{2}}$$</p>\n\n<p>The 2-to-1 map from the torus to the sphere is then $\\psi_0^{-1} H \\phi_1$, where $\\psi_0^{-1}$ and $H$ are as previously defined.</p>\n" }, { "answer_id": 260781, "score": 2, "text": "<p>I hazarded this guess by</p>\n\n<ul>\n<li>looking at some crude pictures and at some published graphs that were almost certainly quite accurate, and</li>\n<li>seeing that the proposition is true in the two trivial cases, which are the $xy$-plane and the $xz$-plane, and</li>\n<li>seeing that it is also true of the planes $x=\\pm z$; this follows from a theorem of geometry published in the 1890s that is stronger than Villarceau's theorem. That stronger theorem says that certain loxodromes of any standard embedded torus lie in a plane.</li>\n</ul>\n\n<p>However yesterday I went over this the old-fashioned way, using some tables that were published in the 19th century. And a calculator for values of the arcsine. The proposition as stated is false, but comes fairly close to the truth numerically. Impressive counterfeits, but detectable if you're careful. Now I'm wondering if a slight modification of the curve that gets revolved about the $z$-axis would make this true.</p>\n" }, { "answer_id": 261263, "score": 12, "text": "<p>This is more comment and question than answer, but I'm writing it as an answer because it's too long to fit as a comment.</p>\n\n<p>Michael Hardy kindly pointed me to a reference on the original mapping between the plane and the sphere:</p>\n\n<p><a href=\"http://www.jstor.org/stable/2369491\" rel=\"noreferrer\">“A Quincuncial Projection of the Sphere”</a>, C. S. Peirce, <em>American Journal of Mathematics</em>, Vol. 2, No. 4 (Dec., 1879), pp. 394-396</p>\n\n<p>Here, Peirce defines the mapping between a point $(x,y)$ in the plane and a point on the sphere with co-latitude $p$ and longitude $\\theta$ as:</p>\n\n<p>$$\\operatorname{cn}\\left(x + i y \\mid \\frac{1}{2}\\right) = \\tan\\left(\\frac{p}{2} \\right)\\left( \\cos(\\theta) + i \\sin(\\theta) \\right)$$</p>\n\n<p>This $\\operatorname{cn}$ function has periods $4 K$ and $2(1+i)K$, where $K=F\\left(\\frac{1}{2}\\right)$. It has zeroes at $(2q+1) K + 2r K i$ and poles at $2q K + (2r+1) K i$, for integers $q, r$. To be clear, the parameters for the Jacobi $\\text{cn}$ function and the complete elliptic integral $F$ are what is usually called $m$, as opposed to the modulus, usually called $k$, where $m=k^2$.</p>\n\n<p>The tables in Peirce are rescaled to make the factor of $K$ go away, but they also seem to be flipped and translated relative to his initial definition. I can reproduce Peirce’s tables with Mathematica if I take them to be describing $x',y'$ where, in relation to his original $x, y$ we have:</p>\n\n<p>$$x'+ i y' = 1 - \\frac{(x + i y)}{K}$$</p>\n\n<p>Now, Michael Hardy requires a function on $(\\mathbb{R}/(2\\pi))^2$, construed as a square in the complex plane, that has periods $2\\pi$ and $2 \\pi i$, zeroes at $0$ and $\\pi+i \\pi$ and poles at $\\pi$ and $i \\pi$. We can write this as:</p>\n\n<p>$$H(z) = \\operatorname{cn}\\left(K \\left(\\left(\\frac{1+i}{\\pi}\\right) z + 1 \\right) \\mid \\frac{1}{2}\\right)$$</p>\n\n<p>He defines a conformal map between a torus in $\\mathbb{R}^3$ and this square in the complex plane via:</p>\n\n<p>$$\\phi_0^{-1}(\\alpha + \\beta i) = \\frac{(\\cos(\\beta),\\sin(\\beta),\\sin(\\alpha))}{\\sqrt{2}-\\cos(\\alpha)}$$</p>\n\n<p>We also have the stereographic map, projecting from the south pole of the unit sphere onto a plane through the equator, whose inverse can be written:</p>\n\n<p>$$\\psi_0^{-1}\\left(\\tan\\left(\\frac{p}{2}\\right)\\left( \\cos(\\theta) + i \\sin(\\theta) \\right)\\right) = (\\cos(\\theta) \\sin(p), \\sin(\\theta) \\sin(p), \\cos(p))$$</p>\n\n<p>[<strong>Edited to add, 21 Feb 2017:</strong> I’ve just belatedly noticed that in the question, the stereographic projection is defined with the poles of the unit sphere on the $x$-axis, rather than the usual convention of putting the poles on the $z$-axis. This explains why I was unable to verify the original claims, and ended up reorienting the torus to make things work.]</p>\n\n<p>Now, the conjecture is that there is a family of planes $\\scr{P}$ such that for any plane $P \\in \\scr{P}$ we have:</p>\n\n<p>$$\\psi^{-1}(H(\\phi(\\scr{T} \\cap P))) = \\scr{S} \\cap P$$</p>\n\n<p>where $\\scr{T}$ is the torus, $\\scr{S}$ is the sphere, and I've dropped the subscripts $0$ from the maps because I want to allow the original versions to be adjusted slightly.</p>\n\n<p>To see how to adjust these maps, note that we want $\\scr{P}$ to include both the $x y$-plane and the $x z$-plane. The stereographic projection of the equator of the sphere is just the unit circle in the complex plane, and the preimage of the unit circle under $H$ consists of the four lines $\\alpha=\\frac{\\pi}{2}$, $\\alpha=\\frac{3\\pi}{2}$, $\\beta=\\frac{\\pi}{2}$ and $\\beta=\\frac{3\\pi}{2}$, where we’re writing points in the fundamental domain as $z=\\alpha + \\beta i$.</p>\n\n<p>The map $\\phi_0^{-1}$ from the fundamental domain to the torus maps the lines $\\alpha=0$ and $\\alpha=\\pi$ to the two circles on the torus that lie in the $x y$-plane. So we need to define:</p>\n\n<p>$$\\phi^{-1}(\\alpha + \\beta i) = \\phi_0^{-1}\\left(\\left(\\alpha + \\frac{\\pi}{2}\\right) + \\beta i\\right) = \\frac{(\\cos(\\beta),\\sin(\\beta),\\cos(\\alpha))}{\\sqrt{2}+\\sin(\\alpha)}$$</p>\n\n<p>To accommodate the $x z$-plane, note that the intersection of the torus with this plane corresponds to the lines $\\beta=0$ and $\\beta=\\pi$, the first of which is mapped by the initial linear transformation in $H$ to a line at 45 degrees that passes through the point $z=K$ in the domain of the original Jacobi function, and contains alternating zeroes and poles of the function. If we take a midpoint between a zero and a pole, such as $\\frac{3}{2}K+\\frac{1}{2}K i$, the Jacobi function evaluates to $-\\sqrt{\\frac{1}{2}}(1+i)$, which corresponds to a longitude of $\\frac{5\\pi}{4}$ on the sphere. We can’t remedy this by offsetting $\\beta$ in the torus map, because the <em>only</em> lines of constant $\\beta$ in the fundamental domain of $H$ that ultimately map to meridians on the sphere are $\\beta=0$ and $\\beta=\\pi$. So we need to redefine the map from the complex plane to the sphere by adding a rotation of $-\\frac{\\pi}{4}$:</p>\n\n<p>$$\\psi^{-1}(z) = \\psi_0^{-1}(\\exp(-i \\pi/4) z)$$</p>\n\n<p>Having made these changes, let’s see exactly what happens for the two planes.</p>\n\n<p>The lines $\\alpha=\\frac{\\pi}{2}$ and $\\alpha=\\frac{3\\pi}{2}$ clearly map under $\\phi^{-1}$ to the two circles in the $x y$-plane where it intersects the torus $\\scr{T}$. Under $H$, they map 2-to-1 to two opposite quadrants of the unit circle (those with arguments between $-\\frac{\\pi}{2}$ and $-\\pi$, and between $0$ and $\\frac{\\pi}{2}$), and then under $\\psi^{-1}$ to two opposite quadrants of the equator. So they don't map to the <em>complete</em> equator; the preimage of the full equator in the fundamental domain of $H$ includes $\\beta=\\frac{\\pi}{2}$ and $\\beta=\\frac{3\\pi}{2}$. But certainly they map into the same plane as the image on the torus.</p>\n\n<p>For the lines $\\beta=0$ and $\\beta=\\pi$, it’s easy to see that for the torus $\\phi^{-1}$ gives us two circles in the $x z$-plane. Under $H$ and $\\psi^{-1}$, the two lines each separately cover the entire circle where the sphere intersects the $x z$-plane. So in this case, the conjecture is exactly true.</p>\n\n<p>However, given the maps that seem to have been forced on us by the need to make things work for these two planes, I can’t see how to make the case $x=\\pm z$ work, or $y =\\pm z$ (I'm not sure whether the claim that the conjecture is true for the former is a typo and the latter was intended, because the initial conjecture was for planes containing the $x$-axis, and $x=\\pm z$ do not contain the $x$-axis).</p>\n\n<p>For the case $y =\\pm z$, consider the curve (in fact, a straight line) in the fundamental domain of $H$ given by:</p>\n\n<p>$$\\alpha = \\cos^{-1}(\\gamma)$$\n$$\\beta = \\sin^{-1}(\\gamma)$$</p>\n\n<p>Since $\\sin(\\beta)$ and $\\cos(\\alpha)$ will be equal, the image of this line under $\\phi^{-1}$ lies in the plane $y=z$.</p>\n\n<p>Suppose we put $\\gamma=\\frac{1}{\\sqrt{2}}$, giving us the point:</p>\n\n<p>$$p_0 = \\frac{(1+i)\\pi}{4}$$</p>\n\n<p>We have:</p>\n\n<p>$$\\phi^{-1}(p_0) = \\left(\\frac{1}{3}, \\frac{1}{3}, \\frac{1}{3}\\right)$$</p>\n\n<p>However:</p>\n\n<p>$$H(p_0) = -\\sqrt{\\sqrt{2}-1} i$$\n$$\\psi^{-1}(H(p_0)) = \\left(-\\sqrt{\\sqrt{2}-1},-\\sqrt{\\sqrt{2}-1},\\sqrt{2}-1\\right)$$</p>\n\n<p>which is not in the plane $y=z$.</p>\n\n<p><strong>Edited to add:</strong></p>\n\n<p>These plots might be helpful. They show the $(\\alpha, \\beta)$ domain, coloured according to which plane the image of each point lies on, when sent either to the sphere or to the torus. Here, I am using the <strong>original</strong> maps, $\\psi_0^{-1} H$ and $\\phi_0^{-1}$ to take the $(\\alpha, \\beta)$ domain to the surfaces of the sphere and the torus. The hue of each point depends on the angle measured around the specified axis in $\\mathbb{R}^3$, modulo $\\pi$, with points coloured black when the angle lies close to a multiple of $\\pi/4$.</p>\n\n<p><a href=\"https://i.stack.imgur.com/wUESs.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/wUESs.png\" alt=\"Planes through (1,0,0)\"></a>\n<a href=\"https://i.stack.imgur.com/1JgIG.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/1JgIG.png\" alt=\"Planes through (0,1,0)\"></a>\n<a href=\"https://i.stack.imgur.com/7UquT.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/7UquT.png\" alt=\"Planes through (0,0,1)\"></a>\n<a href=\"https://i.stack.imgur.com/og2bC.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/og2bC.png\" alt=\"Planes through (1,1,0)\"></a>\n<a href=\"https://i.stack.imgur.com/rRqNh.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/rRqNh.png\" alt=\"Planes through (1,0,1)\"></a>\n<a href=\"https://i.stack.imgur.com/LaHKE.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/LaHKE.png\" alt=\"Planes through (0,1,1)\"></a></p>\n\n<p><strong>Edited to add:</strong> Noting the similarity between the plot for the torus with planes around the $x$-axis and the plot for the sphere with planes around the $z$-axis, I've found that the two can be brought into very close alignment by a change of basis for the torus that preserves the axis, but measures the angles in the opposite direction and translates them all by $\\pi/4$. To be precise, if we use the standard basis for $\\mathbb{R}^3$ for the sphere, and the basis:</p>\n\n<p>$$\\left\\{\\left(0,\\frac{1}{\\sqrt{2}},-\\frac{1}{\\sqrt{2}}\\right),\\left(0,-\\frac{1}{\\sqrt{2}},-\\frac{1}{\\sqrt{2}}\\right),\\left(-1,0,0\\right)\\right\\}$$</p>\n\n<p>for the torus, and then compare the azimuthal angles in polar coordinates adapted to the respective bases, then these angles are very similar, differing from each other by an average of about 0.025.</p>\n\n<p>And the angles agree <strong>exactly</strong> when they are multiples of $\\pi/4$.</p>\n\n<p>I guess this must correspond to Michael Hardy's original observation.</p>\n\n<p><strong>Edited to add:</strong> This is what the final map from the torus to the sphere looks like:</p>\n\n<p><a href=\"https://i.stack.imgur.com/rE5UT.png\" rel=\"noreferrer\"><img src=\"https://i.stack.imgur.com/rE5UT.png\" alt=\"Map from torus to sphere\"></a></p>\n\n<p>Meridians pulled back from the sphere are in black, lines of latitude in red. The pullbacks of the meridians come very close to lying in the same planes as the original meridians on the sphere, if we define:</p>\n\n<p>$$\\phi_1^{-1}(\\alpha + \\beta i) = \\frac{(\\frac{1}{\\sqrt{2}}(\\sin(\\alpha)-\\sin(\\beta)),\\frac{1}{\\sqrt{2}}(\\sin(\\alpha)+\\sin(\\beta)),\\cos(\\beta))}{\\cos(\\alpha)-\\sqrt{2}}$$</p>\n\n<p>The 2-to-1 map from the torus to the sphere is then $\\psi_0^{-1} H \\phi_1$, where $\\psi_0^{-1}$ and $H$ are as previously defined.</p>\n" }, { "answer_id": 260781, "score": 2, "text": "<p>I hazarded this guess by</p>\n\n<ul>\n<li>looking at some crude pictures and at some published graphs that were almost certainly quite accurate, and</li>\n<li>seeing that the proposition is true in the two trivial cases, which are the $xy$-plane and the $xz$-plane, and</li>\n<li>seeing that it is also true of the planes $x=\\pm z$; this follows from a theorem of geometry published in the 1890s that is stronger than Villarceau's theorem. That stronger theorem says that certain loxodromes of any standard embedded torus lie in a plane.</li>\n</ul>\n\n<p>However yesterday I went over this the old-fashioned way, using some tables that were published in the 19th century. And a calculator for values of the arcsine. The proposition as stated is false, but comes fairly close to the truth numerically. Impressive counterfeits, but detectable if you're careful. Now I'm wondering if a slight modification of the curve that gets revolved about the $z$-axis would make this true.</p>\n" } ]
Question: Jacobi&#39;s elliptic functions and plane sections of a torus Answer 1: <p>This is more comment and question than answer, but I'm writing it as an answer because it's too long to fit as a comment.</p> <p>Michael Hardy kindly pointed me to a reference on the original mapping between the plane and the sphere:</p> <p><a href="http://www.jstor.org/stable/2369491" rel="noreferrer">“A Quincuncial Projection of the Sphere”</a>, C. S. Peirce, <em>American Journal of Mathematics</em>, Vol. 2, No. 4 (Dec., 1879), pp. 394-396</p> <p>Here, Peirce defines the mapping between a point $(x,y)$ in the plane and a point on the sphere with co-latitude $p$ and longitude $\theta$ as:</p> <p>$$\operatorname{cn}\left(x + i y \mid \frac{1}{2}\right) = \tan\left(\frac{p}{2} \right)\left( \cos(\theta) + i \sin(\theta) \right)$$</p> <p>This $\operatorname{cn}$ function has periods $4 K$ and $2(1+i)K$, where $K=F\left(\frac{1}{2}\right)$. It has zeroes at $(2q+1) K + 2r K i$ and poles at $2q K + (2r+1) K i$, for integers $q, r$. To be clear, the parameters for the Jacobi $\text{cn}$ function and the complete elliptic integral $F$ are what is usually called $m$, as opposed to the modulus, usually called $k$, where $m=k^2$.</p> <p>The tables in Peirce are rescaled to make the factor of $K$ go away, but they also seem to be flipped and translated relative to his initial definition. I can reproduce Peirce’s tables with Mathematica if I take them to be describing $x',y'$ where, in relation to his original $x, y$ we have:</p> <p>$$x'+ i y' = 1 - \frac{(x + i y)}{K}$$</p> <p>Now, Michael Hardy requires a function on $(\mathbb{R}/(2\pi))^2$, construed as a square in the complex plane, that has periods $2\pi$ and $2 \pi i$, zeroes at $0$ and $\pi+i \pi$ and poles at $\pi$ and $i \pi$. We can write this as:</p> <p>$$H(z) = \operatorname{cn}\left(K \left(\left(\frac{1+i}{\pi}\right) z + 1 \right) \mid \frac{1}{2}\right)$$</p> <p>He defines a conformal map between a torus in $\mathbb{R}^3$ and this square in the complex plane via:</p> <p>$$\phi_0^{-1}(\alpha + \beta i) = \frac{(\cos(\beta),\sin(\beta),\sin(\alpha))}{\sqrt{2}-\cos(\alpha)}$$</p> <p>We also have the stereographic map, projecting from the south pole of the unit sphere onto a plane through the equator, whose inverse can be written:</p> <p>$$\psi_0^{-1}\left(\tan\left(\frac{p}{2}\right)\left( \cos(\theta) + i \sin(\theta) \right)\right) = (\cos(\theta) \sin(p), \sin(\theta) \sin(p), \cos(p))$$</p> <p>[<strong>Edited to add, 21 Feb 2017:</strong> I’ve just belatedly noticed that in the question, the stereographic projection is defined with the poles of the unit sphere on the $x$-axis, rather than the usual convention of putting the poles on the $z$-axis. This explains why I was unable to verify the original claims, and ended up reorienting the torus to make things work.]</p> <p>Now, the conjecture is that there is a family of planes $\scr{P}$ such that for any plane $P \in \scr{P}$ we have:</p> <p>$$\psi^{-1}(H(\phi(\scr{T} \cap P))) = \scr{S} \cap P$$</p> <p>where $\scr{T}$ is the torus, $\scr{S}$ is the sphere, and I've dropped the subscripts $0$ from the maps because I want to allow the original versions to be adjusted slightly.</p> <p>To see how to adjust these maps, note that we want $\scr{P}$ to include both the $x y$-plane and the $x z$-plane. The stereographic projection of the equator of the sphere is just the unit circle in the complex plane, and the preimage of the unit circle under $H$ consists of the four lines $\alpha=\frac{\pi}{2}$, $\alpha=\frac{3\pi}{2}$, $\beta=\frac{\pi}{2}$ and $\beta=\frac{3\pi}{2}$, where we’re writing points in the fundamental domain as $z=\alpha + \beta i$.</p> <p>The map $\phi_0^{-1}$ from the fundamental domain to the torus maps the lines $\alpha=0$ and $\alpha=\pi$ to the two circles on the torus that lie in the $x y$-plane. So we need to define:</p> <p>$$\phi^{-1}(\alpha + \beta i) = \phi_0^{-1}\left(\left(\alpha + \frac{\pi}{2}\right) + \beta i\right) = \frac{(\cos(\beta),\sin(\beta),\cos(\alpha))}{\sqrt{2}+\sin(\alpha)}$$</p> <p>To accommodate the $x z$-plane, note that the intersection of the torus with this plane corresponds to the lines $\beta=0$ and $\beta=\pi$, the first of which is mapped by the initial linear transformation in $H$ to a line at 45 degrees that passes through the point $z=K$ in the domain of the original Jacobi function, and contains alternating zeroes and poles of the function. If we take a midpoint between a zero and a pole, such as $\frac{3}{2}K+\frac{1}{2}K i$, the Jacobi function evaluates to $-\sqrt{\frac{1}{2}}(1+i)$, which corresponds to a longitude of $\frac{5\pi}{4}$ on the sphere. We can’t remedy this by offsetting $\beta$ in the torus map, because the <em>only</em> lines of constant $\beta$ in the fundamental domain of $H$ that ultimately map to meridians on the sphere are $\beta=0$ and $\beta=\pi$. So we need to redefine the map from the complex plane to the sphere by adding a rotation of $-\frac{\pi}{4}$:</p> <p>$$\psi^{-1}(z) = \psi_0^{-1}(\exp(-i \pi/4) z)$$</p> <p>Having made these changes, let’s see exactly what happens for the two planes.</p> <p>The lines $\alpha=\frac{\pi}{2}$ and $\alpha=\frac{3\pi}{2}$ clearly map under $\phi^{-1}$ to the two circles in the $x y$-plane where it intersects the torus $\scr{T}$. Under $H$, they map 2-to-1 to two opposite quadrants of the unit circle (those with arguments between $-\frac{\pi}{2}$ and $-\pi$, and between $0$ and $\frac{\pi}{2}$), and then under $\psi^{-1}$ to two opposite quadrants of the equator. So they don't map to the <em>complete</em> equator; the preimage of the full equator in the fundamental domain of $H$ includes $\beta=\frac{\pi}{2}$ and $\beta=\frac{3\pi}{2}$. But certainly they map into the same plane as the image on the torus.</p> <p>For the lines $\beta=0$ and $\beta=\pi$, it’s easy to see that for the torus $\phi^{-1}$ gives us two circles in the $x z$-plane. Under $H$ and $\psi^{-1}$, the two lines each separately cover the entire circle where the sphere intersects the $x z$-plane. So in this case, the conjecture is exactly true.</p> <p>However, given the maps that seem to have been forced on us by the need to make things work for these two planes, I can’t see how to make the case $x=\pm z$ work, or $y =\pm z$ (I'm not sure whether the claim that the conjecture is true for the former is a typo and the latter was intended, because the initial conjecture was for planes containing the $x$-axis, and $x=\pm z$ do not contain the $x$-axis).</p> <p>For the case $y =\pm z$, consider the curve (in fact, a straight line) in the fundamental domain of $H$ given by:</p> <p>$$\alpha = \cos^{-1}(\gamma)$$ $$\beta = \sin^{-1}(\gamma)$$</p> <p>Since $\sin(\beta)$ and $\cos(\alpha)$ will be equal, the image of this line under $\phi^{-1}$ lies in the plane $y=z$.</p> <p>Suppose we put $\gamma=\frac{1}{\sqrt{2}}$, giving us the point:</p> <p>$$p_0 = \frac{(1+i)\pi}{4}$$</p> <p>We have:</p> <p>$$\phi^{-1}(p_0) = \left(\frac{1}{3}, \frac{1}{3}, \frac{1}{3}\right)$$</p> <p>However:</p> <p>$$H(p_0) = -\sqrt{\sqrt{2}-1} i$$ $$\psi^{-1}(H(p_0)) = \left(-\sqrt{\sqrt{2}-1},-\sqrt{\sqrt{2}-1},\sqrt{2}-1\right)$$</p> <p>which is not in the plane $y=z$.</p> <p><strong>Edited to add:</strong></p> <p>These plots might be helpful. They show the $(\alpha, \beta)$ domain, coloured according to which plane the image of each point lies on, when sent either to the sphere or to the torus. Here, I am using the <strong>original</strong> maps, $\psi_0^{-1} H$ and $\phi_0^{-1}$ to take the $(\alpha, \beta)$ domain to the surfaces of the sphere and the torus. The hue of each point depends on the angle measured around the specified axis in $\mathbb{R}^3$, modulo $\pi$, with points coloured black when the angle lies close to a multiple of $\pi/4$.</p> <p><a href="https://i.stack.imgur.com/wUESs.png" rel="noreferrer"><img src="https://i.stack.imgur.com/wUESs.png" alt="Planes through (1,0,0)"></a> <a href="https://i.stack.imgur.com/1JgIG.png" rel="noreferrer"><img src="https://i.stack.imgur.com/1JgIG.png" alt="Planes through (0,1,0)"></a> <a href="https://i.stack.imgur.com/7UquT.png" rel="noreferrer"><img src="https://i.stack.imgur.com/7UquT.png" alt="Planes through (0,0,1)"></a> <a href="https://i.stack.imgur.com/og2bC.png" rel="noreferrer"><img src="https://i.stack.imgur.com/og2bC.png" alt="Planes through (1,1,0)"></a> <a href="https://i.stack.imgur.com/rRqNh.png" rel="noreferrer"><img src="https://i.stack.imgur.com/rRqNh.png" alt="Planes through (1,0,1)"></a> <a href="https://i.stack.imgur.com/LaHKE.png" rel="noreferrer"><img src="https://i.stack.imgur.com/LaHKE.png" alt="Planes through (0,1,1)"></a></p> <p><strong>Edited to add:</strong> Noting the similarity between the plot for the torus with planes around the $x$-axis and the plot for the sphere with planes around the $z$-axis, I've found that the two can be brought into very close alignment by a change of basis for the torus that preserves the axis, but measures the angles in the opposite direction and translates them all by $\pi/4$. To be precise, if we use the standard basis for $\mathbb{R}^3$ for the sphere, and the basis:</p> <p>$$\left\{\left(0,\frac{1}{\sqrt{2}},-\frac{1}{\sqrt{2}}\right),\left(0,-\frac{1}{\sqrt{2}},-\frac{1}{\sqrt{2}}\right),\left(-1,0,0\right)\right\}$$</p> <p>for the torus, and then compare the azimuthal angles in polar coordinates adapted to the respective bases, then these angles are very similar, differing from each other by an average of about 0.025.</p> <p>And the angles agree <strong>exactly</strong> when they are multiples of $\pi/4$.</p> <p>I guess this must correspond to Michael Hardy's original observation.</p> <p><strong>Edited to add:</strong> This is what the final map from the torus to the sphere looks like:</p> <p><a href="https://i.stack.imgur.com/rE5UT.png" rel="noreferrer"><img src="https://i.stack.imgur.com/rE5UT.png" alt="Map from torus to sphere"></a></p> <p>Meridians pulled back from the sphere are in black, lines of latitude in red. The pullbacks of the meridians come very close to lying in the same planes as the original meridians on the sphere, if we define:</p> <p>$$\phi_1^{-1}(\alpha + \beta i) = \frac{(\frac{1}{\sqrt{2}}(\sin(\alpha)-\sin(\beta)),\frac{1}{\sqrt{2}}(\sin(\alpha)+\sin(\beta)),\cos(\beta))}{\cos(\alpha)-\sqrt{2}}$$</p> <p>The 2-to-1 map from the torus to the sphere is then $\psi_0^{-1} H \phi_1$, where $\psi_0^{-1}$ and $H$ are as previously defined.</p> Answer 2: <p>I hazarded this guess by</p> <ul> <li>looking at some crude pictures and at some published graphs that were almost certainly quite accurate, and</li> <li>seeing that the proposition is true in the two trivial cases, which are the $xy$-plane and the $xz$-plane, and</li> <li>seeing that it is also true of the planes $x=\pm z$; this follows from a theorem of geometry published in the 1890s that is stronger than Villarceau's theorem. That stronger theorem says that certain loxodromes of any standard embedded torus lie in a plane.</li> </ul> <p>However yesterday I went over this the old-fashioned way, using some tables that were published in the 19th century. And a calculator for values of the arcsine. The proposition as stated is false, but comes fairly close to the truth numerically. Impressive counterfeits, but detectable if you're careful. Now I'm wondering if a slight modification of the curve that gets revolved about the $z$-axis would make this true.</p> Answer 3: <p>This is more comment and question than answer, but I'm writing it as an answer because it's too long to fit as a comment.</p> <p>Michael Hardy kindly pointed me to a reference on the original mapping between the plane and the sphere:</p> <p><a href="http://www.jstor.org/stable/2369491" rel="noreferrer">“A Quincuncial Projection of the Sphere”</a>, C. S. Peirce, <em>American Journal of Mathematics</em>, Vol. 2, No. 4 (Dec., 1879), pp. 394-396</p> <p>Here, Peirce defines the mapping between a point $(x,y)$ in the plane and a point on the sphere with co-latitude $p$ and longitude $\theta$ as:</p> <p>$$\operatorname{cn}\left(x + i y \mid \frac{1}{2}\right) = \tan\left(\frac{p}{2} \right)\left( \cos(\theta) + i \sin(\theta) \right)$$</p> <p>This $\operatorname{cn}$ function has periods $4 K$ and $2(1+i)K$, where $K=F\left(\frac{1}{2}\right)$. It has zeroes at $(2q+1) K + 2r K i$ and poles at $2q K + (2r+1) K i$, for integers $q, r$. To be clear, the parameters for the Jacobi $\text{cn}$ function and the complete elliptic integral $F$ are what is usually called $m$, as opposed to the modulus, usually called $k$, where $m=k^2$.</p> <p>The tables in Peirce are rescaled to make the factor of $K$ go away, but they also seem to be flipped and translated relative to his initial definition. I can reproduce Peirce’s tables with Mathematica if I take them to be describing $x',y'$ where, in relation to his original $x, y$ we have:</p> <p>$$x'+ i y' = 1 - \frac{(x + i y)}{K}$$</p> <p>Now, Michael Hardy requires a function on $(\mathbb{R}/(2\pi))^2$, construed as a square in the complex plane, that has periods $2\pi$ and $2 \pi i$, zeroes at $0$ and $\pi+i \pi$ and poles at $\pi$ and $i \pi$. We can write this as:</p> <p>$$H(z) = \operatorname{cn}\left(K \left(\left(\frac{1+i}{\pi}\right) z + 1 \right) \mid \frac{1}{2}\right)$$</p> <p>He defines a conformal map between a torus in $\mathbb{R}^3$ and this square in the complex plane via:</p> <p>$$\phi_0^{-1}(\alpha + \beta i) = \frac{(\cos(\beta),\sin(\beta),\sin(\alpha))}{\sqrt{2}-\cos(\alpha)}$$</p> <p>We also have the stereographic map, projecting from the south pole of the unit sphere onto a plane through the equator, whose inverse can be written:</p> <p>$$\psi_0^{-1}\left(\tan\left(\frac{p}{2}\right)\left( \cos(\theta) + i \sin(\theta) \right)\right) = (\cos(\theta) \sin(p), \sin(\theta) \sin(p), \cos(p))$$</p> <p>[<strong>Edited to add, 21 Feb 2017:</strong> I’ve just belatedly noticed that in the question, the stereographic projection is defined with the poles of the unit sphere on the $x$-axis, rather than the usual convention of putting the poles on the $z$-axis. This explains why I was unable to verify the original claims, and ended up reorienting the torus to make things work.]</p> <p>Now, the conjecture is that there is a family of planes $\scr{P}$ such that for any plane $P \in \scr{P}$ we have:</p> <p>$$\psi^{-1}(H(\phi(\scr{T} \cap P))) = \scr{S} \cap P$$</p> <p>where $\scr{T}$ is the torus, $\scr{S}$ is the sphere, and I've dropped the subscripts $0$ from the maps because I want to allow the original versions to be adjusted slightly.</p> <p>To see how to adjust these maps, note that we want $\scr{P}$ to include both the $x y$-plane and the $x z$-plane. The stereographic projection of the equator of the sphere is just the unit circle in the complex plane, and the preimage of the unit circle under $H$ consists of the four lines $\alpha=\frac{\pi}{2}$, $\alpha=\frac{3\pi}{2}$, $\beta=\frac{\pi}{2}$ and $\beta=\frac{3\pi}{2}$, where we’re writing points in the fundamental domain as $z=\alpha + \beta i$.</p> <p>The map $\phi_0^{-1}$ from the fundamental domain to the torus maps the lines $\alpha=0$ and $\alpha=\pi$ to the two circles on the torus that lie in the $x y$-plane. So we need to define:</p> <p>$$\phi^{-1}(\alpha + \beta i) = \phi_0^{-1}\left(\left(\alpha + \frac{\pi}{2}\right) + \beta i\right) = \frac{(\cos(\beta),\sin(\beta),\cos(\alpha))}{\sqrt{2}+\sin(\alpha)}$$</p> <p>To accommodate the $x z$-plane, note that the intersection of the torus with this plane corresponds to the lines $\beta=0$ and $\beta=\pi$, the first of which is mapped by the initial linear transformation in $H$ to a line at 45 degrees that passes through the point $z=K$ in the domain of the original Jacobi function, and contains alternating zeroes and poles of the function. If we take a midpoint between a zero and a pole, such as $\frac{3}{2}K+\frac{1}{2}K i$, the Jacobi function evaluates to $-\sqrt{\frac{1}{2}}(1+i)$, which corresponds to a longitude of $\frac{5\pi}{4}$ on the sphere. We can’t remedy this by offsetting $\beta$ in the torus map, because the <em>only</em> lines of constant $\beta$ in the fundamental domain of $H$ that ultimately map to meridians on the sphere are $\beta=0$ and $\beta=\pi$. So we need to redefine the map from the complex plane to the sphere by adding a rotation of $-\frac{\pi}{4}$:</p> <p>$$\psi^{-1}(z) = \psi_0^{-1}(\exp(-i \pi/4) z)$$</p> <p>Having made these changes, let’s see exactly what happens for the two planes.</p> <p>The lines $\alpha=\frac{\pi}{2}$ and $\alpha=\frac{3\pi}{2}$ clearly map under $\phi^{-1}$ to the two circles in the $x y$-plane where it intersects the torus $\scr{T}$. Under $H$, they map 2-to-1 to two opposite quadrants of the unit circle (those with arguments between $-\frac{\pi}{2}$ and $-\pi$, and between $0$ and $\frac{\pi}{2}$), and then under $\psi^{-1}$ to two opposite quadrants of the equator. So they don't map to the <em>complete</em> equator; the preimage of the full equator in the fundamental domain of $H$ includes $\beta=\frac{\pi}{2}$ and $\beta=\frac{3\pi}{2}$. But certainly they map into the same plane as the image on the torus.</p> <p>For the lines $\beta=0$ and $\beta=\pi$, it’s easy to see that for the torus $\phi^{-1}$ gives us two circles in the $x z$-plane. Under $H$ and $\psi^{-1}$, the two lines each separately cover the entire circle where the sphere intersects the $x z$-plane. So in this case, the conjecture is exactly true.</p> <p>However, given the maps that seem to have been forced on us by the need to make things work for these two planes, I can’t see how to make the case $x=\pm z$ work, or $y =\pm z$ (I'm not sure whether the claim that the conjecture is true for the former is a typo and the latter was intended, because the initial conjecture was for planes containing the $x$-axis, and $x=\pm z$ do not contain the $x$-axis).</p> <p>For the case $y =\pm z$, consider the curve (in fact, a straight line) in the fundamental domain of $H$ given by:</p> <p>$$\alpha = \cos^{-1}(\gamma)$$ $$\beta = \sin^{-1}(\gamma)$$</p> <p>Since $\sin(\beta)$ and $\cos(\alpha)$ will be equal, the image of this line under $\phi^{-1}$ lies in the plane $y=z$.</p> <p>Suppose we put $\gamma=\frac{1}{\sqrt{2}}$, giving us the point:</p> <p>$$p_0 = \frac{(1+i)\pi}{4}$$</p> <p>We have:</p> <p>$$\phi^{-1}(p_0) = \left(\frac{1}{3}, \frac{1}{3}, \frac{1}{3}\right)$$</p> <p>However:</p> <p>$$H(p_0) = -\sqrt{\sqrt{2}-1} i$$ $$\psi^{-1}(H(p_0)) = \left(-\sqrt{\sqrt{2}-1},-\sqrt{\sqrt{2}-1},\sqrt{2}-1\right)$$</p> <p>which is not in the plane $y=z$.</p> <p><strong>Edited to add:</strong></p> <p>These plots might be helpful. They show the $(\alpha, \beta)$ domain, coloured according to which plane the image of each point lies on, when sent either to the sphere or to the torus. Here, I am using the <strong>original</strong> maps, $\psi_0^{-1} H$ and $\phi_0^{-1}$ to take the $(\alpha, \beta)$ domain to the surfaces of the sphere and the torus. The hue of each point depends on the angle measured around the specified axis in $\mathbb{R}^3$, modulo $\pi$, with points coloured black when the angle lies close to a multiple of $\pi/4$.</p> <p><a href="https://i.stack.imgur.com/wUESs.png" rel="noreferrer"><img src="https://i.stack.imgur.com/wUESs.png" alt="Planes through (1,0,0)"></a> <a href="https://i.stack.imgur.com/1JgIG.png" rel="noreferrer"><img src="https://i.stack.imgur.com/1JgIG.png" alt="Planes through (0,1,0)"></a> <a href="https://i.stack.imgur.com/7UquT.png" rel="noreferrer"><img src="https://i.stack.imgur.com/7UquT.png" alt="Planes through (0,0,1)"></a> <a href="https://i.stack.imgur.com/og2bC.png" rel="noreferrer"><img src="https://i.stack.imgur.com/og2bC.png" alt="Planes through (1,1,0)"></a> <a href="https://i.stack.imgur.com/rRqNh.png" rel="noreferrer"><img src="https://i.stack.imgur.com/rRqNh.png" alt="Planes through (1,0,1)"></a> <a href="https://i.stack.imgur.com/LaHKE.png" rel="noreferrer"><img src="https://i.stack.imgur.com/LaHKE.png" alt="Planes through (0,1,1)"></a></p> <p><strong>Edited to add:</strong> Noting the similarity between the plot for the torus with planes around the $x$-axis and the plot for the sphere with planes around the $z$-axis, I've found that the two can be brought into very close alignment by a change of basis for the torus that preserves the axis, but measures the angles in the opposite direction and translates them all by $\pi/4$. To be precise, if we use the standard basis for $\mathbb{R}^3$ for the sphere, and the basis:</p> <p>$$\left\{\left(0,\frac{1}{\sqrt{2}},-\frac{1}{\sqrt{2}}\right),\left(0,-\frac{1}{\sqrt{2}},-\frac{1}{\sqrt{2}}\right),\left(-1,0,0\right)\right\}$$</p> <p>for the torus, and then compare the azimuthal angles in polar coordinates adapted to the respective bases, then these angles are very similar, differing from each other by an average of about 0.025.</p> <p>And the angles agree <strong>exactly</strong> when they are multiples of $\pi/4$.</p> <p>I guess this must correspond to Michael Hardy's original observation.</p> <p><strong>Edited to add:</strong> This is what the final map from the torus to the sphere looks like:</p> <p><a href="https://i.stack.imgur.com/rE5UT.png" rel="noreferrer"><img src="https://i.stack.imgur.com/rE5UT.png" alt="Map from torus to sphere"></a></p> <p>Meridians pulled back from the sphere are in black, lines of latitude in red. The pullbacks of the meridians come very close to lying in the same planes as the original meridians on the sphere, if we define:</p> <p>$$\phi_1^{-1}(\alpha + \beta i) = \frac{(\frac{1}{\sqrt{2}}(\sin(\alpha)-\sin(\beta)),\frac{1}{\sqrt{2}}(\sin(\alpha)+\sin(\beta)),\cos(\beta))}{\cos(\alpha)-\sqrt{2}}$$</p> <p>The 2-to-1 map from the torus to the sphere is then $\psi_0^{-1} H \phi_1$, where $\psi_0^{-1}$ and $H$ are as previously defined.</p> Answer 4: <p>I hazarded this guess by</p> <ul> <li>looking at some crude pictures and at some published graphs that were almost certainly quite accurate, and</li> <li>seeing that the proposition is true in the two trivial cases, which are the $xy$-plane and the $xz$-plane, and</li> <li>seeing that it is also true of the planes $x=\pm z$; this follows from a theorem of geometry published in the 1890s that is stronger than Villarceau's theorem. That stronger theorem says that certain loxodromes of any standard embedded torus lie in a plane.</li> </ul> <p>However yesterday I went over this the old-fashioned way, using some tables that were published in the 19th century. And a calculator for values of the arcsine. The proposition as stated is false, but comes fairly close to the truth numerically. Impressive counterfeits, but detectable if you're careful. Now I'm wondering if a slight modification of the curve that gets revolved about the $z$-axis would make this true.</p> Answer 5: <p>This is more comment and question than answer, but I'm writing it as an answer because it's too long to fit as a comment.</p> <p>Michael Hardy kindly pointed me to a reference on the original mapping between the plane and the sphere:</p> <p><a href="http://www.jstor.org/stable/2369491" rel="noreferrer">“A Quincuncial Projection of the Sphere”</a>, C. S. Peirce, <em>American Journal of Mathematics</em>, Vol. 2, No. 4 (Dec., 1879), pp. 394-396</p> <p>Here, Peirce defines the mapping between a point $(x,y)$ in the plane and a point on the sphere with co-latitude $p$ and longitude $\theta$ as:</p> <p>$$\operatorname{cn}\left(x + i y \mid \frac{1}{2}\right) = \tan\left(\frac{p}{2} \right)\left( \cos(\theta) + i \sin(\theta) \right)$$</p> <p>This $\operatorname{cn}$ function has periods $4 K$ and $2(1+i)K$, where $K=F\left(\frac{1}{2}\right)$. It has zeroes at $(2q+1) K + 2r K i$ and poles at $2q K + (2r+1) K i$, for integers $q, r$. To be clear, the parameters for the Jacobi $\text{cn}$ function and the complete elliptic integral $F$ are what is usually called $m$, as opposed to the modulus, usually called $k$, where $m=k^2$.</p> <p>The tables in Peirce are rescaled to make the factor of $K$ go away, but they also seem to be flipped and translated relative to his initial definition. I can reproduce Peirce’s tables with Mathematica if I take them to be describing $x',y'$ where, in relation to his original $x, y$ we have:</p> <p>$$x'+ i y' = 1 - \frac{(x + i y)}{K}$$</p> <p>Now, Michael Hardy requires a function on $(\mathbb{R}/(2\pi))^2$, construed as a square in the complex plane, that has periods $2\pi$ and $2 \pi i$, zeroes at $0$ and $\pi+i \pi$ and poles at $\pi$ and $i \pi$. We can write this as:</p> <p>$$H(z) = \operatorname{cn}\left(K \left(\left(\frac{1+i}{\pi}\right) z + 1 \right) \mid \frac{1}{2}\right)$$</p> <p>He defines a conformal map between a torus in $\mathbb{R}^3$ and this square in the complex plane via:</p> <p>$$\phi_0^{-1}(\alpha + \beta i) = \frac{(\cos(\beta),\sin(\beta),\sin(\alpha))}{\sqrt{2}-\cos(\alpha)}$$</p> <p>We also have the stereographic map, projecting from the south pole of the unit sphere onto a plane through the equator, whose inverse can be written:</p> <p>$$\psi_0^{-1}\left(\tan\left(\frac{p}{2}\right)\left( \cos(\theta) + i \sin(\theta) \right)\right) = (\cos(\theta) \sin(p), \sin(\theta) \sin(p), \cos(p))$$</p> <p>[<strong>Edited to add, 21 Feb 2017:</strong> I’ve just belatedly noticed that in the question, the stereographic projection is defined with the poles of the unit sphere on the $x$-axis, rather than the usual convention of putting the poles on the $z$-axis. This explains why I was unable to verify the original claims, and ended up reorienting the torus to make things work.]</p> <p>Now, the conjecture is that there is a family of planes $\scr{P}$ such that for any plane $P \in \scr{P}$ we have:</p> <p>$$\psi^{-1}(H(\phi(\scr{T} \cap P))) = \scr{S} \cap P$$</p> <p>where $\scr{T}$ is the torus, $\scr{S}$ is the sphere, and I've dropped the subscripts $0$ from the maps because I want to allow the original versions to be adjusted slightly.</p> <p>To see how to adjust these maps, note that we want $\scr{P}$ to include both the $x y$-plane and the $x z$-plane. The stereographic projection of the equator of the sphere is just the unit circle in the complex plane, and the preimage of the unit circle under $H$ consists of the four lines $\alpha=\frac{\pi}{2}$, $\alpha=\frac{3\pi}{2}$, $\beta=\frac{\pi}{2}$ and $\beta=\frac{3\pi}{2}$, where we’re writing points in the fundamental domain as $z=\alpha + \beta i$.</p> <p>The map $\phi_0^{-1}$ from the fundamental domain to the torus maps the lines $\alpha=0$ and $\alpha=\pi$ to the two circles on the torus that lie in the $x y$-plane. So we need to define:</p> <p>$$\phi^{-1}(\alpha + \beta i) = \phi_0^{-1}\left(\left(\alpha + \frac{\pi}{2}\right) + \beta i\right) = \frac{(\cos(\beta),\sin(\beta),\cos(\alpha))}{\sqrt{2}+\sin(\alpha)}$$</p> <p>To accommodate the $x z$-plane, note that the intersection of the torus with this plane corresponds to the lines $\beta=0$ and $\beta=\pi$, the first of which is mapped by the initial linear transformation in $H$ to a line at 45 degrees that passes through the point $z=K$ in the domain of the original Jacobi function, and contains alternating zeroes and poles of the function. If we take a midpoint between a zero and a pole, such as $\frac{3}{2}K+\frac{1}{2}K i$, the Jacobi function evaluates to $-\sqrt{\frac{1}{2}}(1+i)$, which corresponds to a longitude of $\frac{5\pi}{4}$ on the sphere. We can’t remedy this by offsetting $\beta$ in the torus map, because the <em>only</em> lines of constant $\beta$ in the fundamental domain of $H$ that ultimately map to meridians on the sphere are $\beta=0$ and $\beta=\pi$. So we need to redefine the map from the complex plane to the sphere by adding a rotation of $-\frac{\pi}{4}$:</p> <p>$$\psi^{-1}(z) = \psi_0^{-1}(\exp(-i \pi/4) z)$$</p> <p>Having made these changes, let’s see exactly what happens for the two planes.</p> <p>The lines $\alpha=\frac{\pi}{2}$ and $\alpha=\frac{3\pi}{2}$ clearly map under $\phi^{-1}$ to the two circles in the $x y$-plane where it intersects the torus $\scr{T}$. Under $H$, they map 2-to-1 to two opposite quadrants of the unit circle (those with arguments between $-\frac{\pi}{2}$ and $-\pi$, and between $0$ and $\frac{\pi}{2}$), and then under $\psi^{-1}$ to two opposite quadrants of the equator. So they don't map to the <em>complete</em> equator; the preimage of the full equator in the fundamental domain of $H$ includes $\beta=\frac{\pi}{2}$ and $\beta=\frac{3\pi}{2}$. But certainly they map into the same plane as the image on the torus.</p> <p>For the lines $\beta=0$ and $\beta=\pi$, it’s easy to see that for the torus $\phi^{-1}$ gives us two circles in the $x z$-plane. Under $H$ and $\psi^{-1}$, the two lines each separately cover the entire circle where the sphere intersects the $x z$-plane. So in this case, the conjecture is exactly true.</p> <p>However, given the maps that seem to have been forced on us by the need to make things work for these two planes, I can’t see how to make the case $x=\pm z$ work, or $y =\pm z$ (I'm not sure whether the claim that the conjecture is true for the former is a typo and the latter was intended, because the initial conjecture was for planes containing the $x$-axis, and $x=\pm z$ do not contain the $x$-axis).</p> <p>For the case $y =\pm z$, consider the curve (in fact, a straight line) in the fundamental domain of $H$ given by:</p> <p>$$\alpha = \cos^{-1}(\gamma)$$ $$\beta = \sin^{-1}(\gamma)$$</p> <p>Since $\sin(\beta)$ and $\cos(\alpha)$ will be equal, the image of this line under $\phi^{-1}$ lies in the plane $y=z$.</p> <p>Suppose we put $\gamma=\frac{1}{\sqrt{2}}$, giving us the point:</p> <p>$$p_0 = \frac{(1+i)\pi}{4}$$</p> <p>We have:</p> <p>$$\phi^{-1}(p_0) = \left(\frac{1}{3}, \frac{1}{3}, \frac{1}{3}\right)$$</p> <p>However:</p> <p>$$H(p_0) = -\sqrt{\sqrt{2}-1} i$$ $$\psi^{-1}(H(p_0)) = \left(-\sqrt{\sqrt{2}-1},-\sqrt{\sqrt{2}-1},\sqrt{2}-1\right)$$</p> <p>which is not in the plane $y=z$.</p> <p><strong>Edited to add:</strong></p> <p>These plots might be helpful. They show the $(\alpha, \beta)$ domain, coloured according to which plane the image of each point lies on, when sent either to the sphere or to the torus. Here, I am using the <strong>original</strong> maps, $\psi_0^{-1} H$ and $\phi_0^{-1}$ to take the $(\alpha, \beta)$ domain to the surfaces of the sphere and the torus. The hue of each point depends on the angle measured around the specified axis in $\mathbb{R}^3$, modulo $\pi$, with points coloured black when the angle lies close to a multiple of $\pi/4$.</p> <p><a href="https://i.stack.imgur.com/wUESs.png" rel="noreferrer"><img src="https://i.stack.imgur.com/wUESs.png" alt="Planes through (1,0,0)"></a> <a href="https://i.stack.imgur.com/1JgIG.png" rel="noreferrer"><img src="https://i.stack.imgur.com/1JgIG.png" alt="Planes through (0,1,0)"></a> <a href="https://i.stack.imgur.com/7UquT.png" rel="noreferrer"><img src="https://i.stack.imgur.com/7UquT.png" alt="Planes through (0,0,1)"></a> <a href="https://i.stack.imgur.com/og2bC.png" rel="noreferrer"><img src="https://i.stack.imgur.com/og2bC.png" alt="Planes through (1,1,0)"></a> <a href="https://i.stack.imgur.com/rRqNh.png" rel="noreferrer"><img src="https://i.stack.imgur.com/rRqNh.png" alt="Planes through (1,0,1)"></a> <a href="https://i.stack.imgur.com/LaHKE.png" rel="noreferrer"><img src="https://i.stack.imgur.com/LaHKE.png" alt="Planes through (0,1,1)"></a></p> <p><strong>Edited to add:</strong> Noting the similarity between the plot for the torus with planes around the $x$-axis and the plot for the sphere with planes around the $z$-axis, I've found that the two can be brought into very close alignment by a change of basis for the torus that preserves the axis, but measures the angles in the opposite direction and translates them all by $\pi/4$. To be precise, if we use the standard basis for $\mathbb{R}^3$ for the sphere, and the basis:</p> <p>$$\left\{\left(0,\frac{1}{\sqrt{2}},-\frac{1}{\sqrt{2}}\right),\left(0,-\frac{1}{\sqrt{2}},-\frac{1}{\sqrt{2}}\right),\left(-1,0,0\right)\right\}$$</p> <p>for the torus, and then compare the azimuthal angles in polar coordinates adapted to the respective bases, then these angles are very similar, differing from each other by an average of about 0.025.</p> <p>And the angles agree <strong>exactly</strong> when they are multiples of $\pi/4$.</p> <p>I guess this must correspond to Michael Hardy's original observation.</p> <p><strong>Edited to add:</strong> This is what the final map from the torus to the sphere looks like:</p> <p><a href="https://i.stack.imgur.com/rE5UT.png" rel="noreferrer"><img src="https://i.stack.imgur.com/rE5UT.png" alt="Map from torus to sphere"></a></p> <p>Meridians pulled back from the sphere are in black, lines of latitude in red. The pullbacks of the meridians come very close to lying in the same planes as the original meridians on the sphere, if we define:</p> <p>$$\phi_1^{-1}(\alpha + \beta i) = \frac{(\frac{1}{\sqrt{2}}(\sin(\alpha)-\sin(\beta)),\frac{1}{\sqrt{2}}(\sin(\alpha)+\sin(\beta)),\cos(\beta))}{\cos(\alpha)-\sqrt{2}}$$</p> <p>The 2-to-1 map from the torus to the sphere is then $\psi_0^{-1} H \phi_1$, where $\psi_0^{-1}$ and $H$ are as previously defined.</p> Answer 6: <p>I hazarded this guess by</p> <ul> <li>looking at some crude pictures and at some published graphs that were almost certainly quite accurate, and</li> <li>seeing that the proposition is true in the two trivial cases, which are the $xy$-plane and the $xz$-plane, and</li> <li>seeing that it is also true of the planes $x=\pm z$; this follows from a theorem of geometry published in the 1890s that is stronger than Villarceau's theorem. That stronger theorem says that certain loxodromes of any standard embedded torus lie in a plane.</li> </ul> <p>However yesterday I went over this the old-fashioned way, using some tables that were published in the 19th century. And a calculator for values of the arcsine. The proposition as stated is false, but comes fairly close to the truth numerically. Impressive counterfeits, but detectable if you're careful. Now I'm wondering if a slight modification of the curve that gets revolved about the $z$-axis would make this true.</p>
Affine spaces as algebras for an operad?
260,487
10
https://mathoverflow.net/questions/260487/affine-spaces-as-algebras-for-an-operad
<p>(at.algebraic-topology because I don't know who else thinks about operads)</p> <p>Let $A$ be an affine space, i.e. a torsor over (the abelian group underlying) a vector space $V$ over a field $K$. Then for any $a_1,\dotsc,a_n \in A$ and any $\lambda_1,\dotsc,\lambda_n\in K$ with $\sum \lambda_i = 1$, we can form the weighted average $$ \lambda_1 a_1 + \dotsb + \lambda_n a_n \in A. $$ I'm trying to work out exactly what structure this exhibits on $A$.</p> <p>Let's let $Pr_K(n) \subset K^n$ denote the subset of <em>probability vectors</em>, i.e. $n$-tuples which sum to 1. I believe this forms an operad: the map $$ Pr_K(n)\times Pr_K(i_1) \times \dotsb \times Pr_K(i_n) \to Pr_K(i_1 + \dotsb + i_n)$$ sends $$\langle \{\lambda_j\}_{j=1}^n; v_1,\dotsc, v_n\rangle \mapsto \lambda_1 v_1 \mathbin| \dotsb \mathbin| \lambda_n v_n$$ where by $|$ I mean "concatenate these tuples". Then I think $A$ should be an algebra over the operad $Pr_K$.</p> <p>Is this correct? Is it written down somewhere? What's the standard name for this operad? Can you do anything with this perspective?</p>
[ { "answer_id": 260500, "score": 6, "text": "<p>In order to present affine spaces in this fashion you need a variation of operads called <em>clones</em> or <em>cartesian operads</em> (which eventually are equivalent to Lawvere theories). You can find some references on <a href=\"https://ncatlab.org/nlab/show/clone\">nLab</a>.</p>\n\n<p>A clone is like a symmetric operad, but with all maps between finite sets acting not just bijections. However, it is more convenient to phrase the definition as follows. A clone is a functor $C \\colon \\mathsf{FinSet} \\to \\mathsf{Set}$ equipped with a family of substitution operators, i.e. for each pair of finite sets $S$ and $T$ we specify a function $(C S)^T \\times C T \\to C S$ and distinguish projection operations $\\pi_s \\in C S$ for each $s \\in S$. These are supposed to satisfy some axioms (see below).</p>\n\n<p>I will call the clone for affine spaces $\\mathbb{A}$. (And I will refrain from using vocabulary from probability theory, this would go better with convex spaces.) For each $S$, $\\mathbb{A} S$ is the set of formal affine combinations of elements of $S$, functoriality of $\\mathbb{A}$ is given by grouping coefficients of points together when they are identified by a function between finite sets. Substitution operations evaluate affine combinations of affine combinations. If I have a combination $b \\in \\mathbb{A} T$ and a family $a \\colon T \\to \\mathbb{A}S$, I will denote the resulting substitution by $a \\bullet b \\in \\mathbb{A}S$. (It is really just a shorthand for $s \\mapsto \\sum_{t \\in T} a_{t, s} b_t$.) Moreover, $\\pi_s = e_s$ is the trivial combination with coefficient $1$ at $s$ and $0$ elsewhere.</p>\n\n<p>We can now give the following definition. An affine space is a set $E$ equipped with operations $E^S \\times \\mathbb{A}S \\to E$, $(x , a) \\mapsto x \\bullet a$ subject to the following axioms.</p>\n\n<ol>\n<li>For every finite set $S$ and all $x \\colon S \\to E$ and $s \\in S$, $x \\bullet e_s = x_s$.</li>\n<li>For every pair of finite sets $S$ and $T$ and all $x \\colon S \\to E$, $a \\colon T \\to \\mathbb{A} S$ and $b \\in \\mathbb{A} T$, $x \\bullet (a \\bullet b) = (x \\bullet a) \\bullet b$.</li>\n<li>For every function between finite sets $\\phi \\colon S \\to S'$ and all $x \\colon S' \\to E$ and $a \\in \\mathbb{A} S$, $(x \\phi) \\bullet a = x \\bullet (\\phi a)$.</li>\n</ol>\n\n<p>In particular, $\\mathbb{A} T$ for a fixed $T$ carries a standard structure of an affine space given by the structure maps of $\\mathbb{A}$. In this case, the axioms specialize to some standard facts about formal affine combinations and they are in fact the same as axioms saying that $\\mathbb{A}$ is a clone in the first place. This way you can recover the full definition of a clone which I neglected to write down above.</p>\n" }, { "answer_id": 260515, "score": 5, "text": "<p>For $K=\\mathbb{R}$, the positive part of your operad (mentioned in Gabriel's comment), and its algebras have been discussed by Tom Leinster and others in connection to entropy. See, for example,</p>\n\n<p><a href=\"https://golem.ph.utexas.edu/category/2011/05/an_operadic_introduction_to_en.html\" rel=\"noreferrer\">https://golem.ph.utexas.edu/category/2011/05/an_operadic_introduction_to_en.html</a></p>\n\n<p><a href=\"http://www.maths.ed.ac.uk/~tl/b.pdf\" rel=\"noreferrer\">http://www.maths.ed.ac.uk/~tl/b.pdf</a></p>\n" }, { "answer_id": 260500, "score": 6, "text": "<p>In order to present affine spaces in this fashion you need a variation of operads called <em>clones</em> or <em>cartesian operads</em> (which eventually are equivalent to Lawvere theories). You can find some references on <a href=\"https://ncatlab.org/nlab/show/clone\">nLab</a>.</p>\n\n<p>A clone is like a symmetric operad, but with all maps between finite sets acting not just bijections. However, it is more convenient to phrase the definition as follows. A clone is a functor $C \\colon \\mathsf{FinSet} \\to \\mathsf{Set}$ equipped with a family of substitution operators, i.e. for each pair of finite sets $S$ and $T$ we specify a function $(C S)^T \\times C T \\to C S$ and distinguish projection operations $\\pi_s \\in C S$ for each $s \\in S$. These are supposed to satisfy some axioms (see below).</p>\n\n<p>I will call the clone for affine spaces $\\mathbb{A}$. (And I will refrain from using vocabulary from probability theory, this would go better with convex spaces.) For each $S$, $\\mathbb{A} S$ is the set of formal affine combinations of elements of $S$, functoriality of $\\mathbb{A}$ is given by grouping coefficients of points together when they are identified by a function between finite sets. Substitution operations evaluate affine combinations of affine combinations. If I have a combination $b \\in \\mathbb{A} T$ and a family $a \\colon T \\to \\mathbb{A}S$, I will denote the resulting substitution by $a \\bullet b \\in \\mathbb{A}S$. (It is really just a shorthand for $s \\mapsto \\sum_{t \\in T} a_{t, s} b_t$.) Moreover, $\\pi_s = e_s$ is the trivial combination with coefficient $1$ at $s$ and $0$ elsewhere.</p>\n\n<p>We can now give the following definition. An affine space is a set $E$ equipped with operations $E^S \\times \\mathbb{A}S \\to E$, $(x , a) \\mapsto x \\bullet a$ subject to the following axioms.</p>\n\n<ol>\n<li>For every finite set $S$ and all $x \\colon S \\to E$ and $s \\in S$, $x \\bullet e_s = x_s$.</li>\n<li>For every pair of finite sets $S$ and $T$ and all $x \\colon S \\to E$, $a \\colon T \\to \\mathbb{A} S$ and $b \\in \\mathbb{A} T$, $x \\bullet (a \\bullet b) = (x \\bullet a) \\bullet b$.</li>\n<li>For every function between finite sets $\\phi \\colon S \\to S'$ and all $x \\colon S' \\to E$ and $a \\in \\mathbb{A} S$, $(x \\phi) \\bullet a = x \\bullet (\\phi a)$.</li>\n</ol>\n\n<p>In particular, $\\mathbb{A} T$ for a fixed $T$ carries a standard structure of an affine space given by the structure maps of $\\mathbb{A}$. In this case, the axioms specialize to some standard facts about formal affine combinations and they are in fact the same as axioms saying that $\\mathbb{A}$ is a clone in the first place. This way you can recover the full definition of a clone which I neglected to write down above.</p>\n" }, { "answer_id": 260515, "score": 5, "text": "<p>For $K=\\mathbb{R}$, the positive part of your operad (mentioned in Gabriel's comment), and its algebras have been discussed by Tom Leinster and others in connection to entropy. See, for example,</p>\n\n<p><a href=\"https://golem.ph.utexas.edu/category/2011/05/an_operadic_introduction_to_en.html\" rel=\"noreferrer\">https://golem.ph.utexas.edu/category/2011/05/an_operadic_introduction_to_en.html</a></p>\n\n<p><a href=\"http://www.maths.ed.ac.uk/~tl/b.pdf\" rel=\"noreferrer\">http://www.maths.ed.ac.uk/~tl/b.pdf</a></p>\n" }, { "answer_id": 260500, "score": 6, "text": "<p>In order to present affine spaces in this fashion you need a variation of operads called <em>clones</em> or <em>cartesian operads</em> (which eventually are equivalent to Lawvere theories). You can find some references on <a href=\"https://ncatlab.org/nlab/show/clone\">nLab</a>.</p>\n\n<p>A clone is like a symmetric operad, but with all maps between finite sets acting not just bijections. However, it is more convenient to phrase the definition as follows. A clone is a functor $C \\colon \\mathsf{FinSet} \\to \\mathsf{Set}$ equipped with a family of substitution operators, i.e. for each pair of finite sets $S$ and $T$ we specify a function $(C S)^T \\times C T \\to C S$ and distinguish projection operations $\\pi_s \\in C S$ for each $s \\in S$. These are supposed to satisfy some axioms (see below).</p>\n\n<p>I will call the clone for affine spaces $\\mathbb{A}$. (And I will refrain from using vocabulary from probability theory, this would go better with convex spaces.) For each $S$, $\\mathbb{A} S$ is the set of formal affine combinations of elements of $S$, functoriality of $\\mathbb{A}$ is given by grouping coefficients of points together when they are identified by a function between finite sets. Substitution operations evaluate affine combinations of affine combinations. If I have a combination $b \\in \\mathbb{A} T$ and a family $a \\colon T \\to \\mathbb{A}S$, I will denote the resulting substitution by $a \\bullet b \\in \\mathbb{A}S$. (It is really just a shorthand for $s \\mapsto \\sum_{t \\in T} a_{t, s} b_t$.) Moreover, $\\pi_s = e_s$ is the trivial combination with coefficient $1$ at $s$ and $0$ elsewhere.</p>\n\n<p>We can now give the following definition. An affine space is a set $E$ equipped with operations $E^S \\times \\mathbb{A}S \\to E$, $(x , a) \\mapsto x \\bullet a$ subject to the following axioms.</p>\n\n<ol>\n<li>For every finite set $S$ and all $x \\colon S \\to E$ and $s \\in S$, $x \\bullet e_s = x_s$.</li>\n<li>For every pair of finite sets $S$ and $T$ and all $x \\colon S \\to E$, $a \\colon T \\to \\mathbb{A} S$ and $b \\in \\mathbb{A} T$, $x \\bullet (a \\bullet b) = (x \\bullet a) \\bullet b$.</li>\n<li>For every function between finite sets $\\phi \\colon S \\to S'$ and all $x \\colon S' \\to E$ and $a \\in \\mathbb{A} S$, $(x \\phi) \\bullet a = x \\bullet (\\phi a)$.</li>\n</ol>\n\n<p>In particular, $\\mathbb{A} T$ for a fixed $T$ carries a standard structure of an affine space given by the structure maps of $\\mathbb{A}$. In this case, the axioms specialize to some standard facts about formal affine combinations and they are in fact the same as axioms saying that $\\mathbb{A}$ is a clone in the first place. This way you can recover the full definition of a clone which I neglected to write down above.</p>\n" }, { "answer_id": 260515, "score": 5, "text": "<p>For $K=\\mathbb{R}$, the positive part of your operad (mentioned in Gabriel's comment), and its algebras have been discussed by Tom Leinster and others in connection to entropy. See, for example,</p>\n\n<p><a href=\"https://golem.ph.utexas.edu/category/2011/05/an_operadic_introduction_to_en.html\" rel=\"noreferrer\">https://golem.ph.utexas.edu/category/2011/05/an_operadic_introduction_to_en.html</a></p>\n\n<p><a href=\"http://www.maths.ed.ac.uk/~tl/b.pdf\" rel=\"noreferrer\">http://www.maths.ed.ac.uk/~tl/b.pdf</a></p>\n" } ]
Question: Affine spaces as algebras for an operad? Answer 1: <p>In order to present affine spaces in this fashion you need a variation of operads called <em>clones</em> or <em>cartesian operads</em> (which eventually are equivalent to Lawvere theories). You can find some references on <a href="https://ncatlab.org/nlab/show/clone">nLab</a>.</p> <p>A clone is like a symmetric operad, but with all maps between finite sets acting not just bijections. However, it is more convenient to phrase the definition as follows. A clone is a functor $C \colon \mathsf{FinSet} \to \mathsf{Set}$ equipped with a family of substitution operators, i.e. for each pair of finite sets $S$ and $T$ we specify a function $(C S)^T \times C T \to C S$ and distinguish projection operations $\pi_s \in C S$ for each $s \in S$. These are supposed to satisfy some axioms (see below).</p> <p>I will call the clone for affine spaces $\mathbb{A}$. (And I will refrain from using vocabulary from probability theory, this would go better with convex spaces.) For each $S$, $\mathbb{A} S$ is the set of formal affine combinations of elements of $S$, functoriality of $\mathbb{A}$ is given by grouping coefficients of points together when they are identified by a function between finite sets. Substitution operations evaluate affine combinations of affine combinations. If I have a combination $b \in \mathbb{A} T$ and a family $a \colon T \to \mathbb{A}S$, I will denote the resulting substitution by $a \bullet b \in \mathbb{A}S$. (It is really just a shorthand for $s \mapsto \sum_{t \in T} a_{t, s} b_t$.) Moreover, $\pi_s = e_s$ is the trivial combination with coefficient $1$ at $s$ and $0$ elsewhere.</p> <p>We can now give the following definition. An affine space is a set $E$ equipped with operations $E^S \times \mathbb{A}S \to E$, $(x , a) \mapsto x \bullet a$ subject to the following axioms.</p> <ol> <li>For every finite set $S$ and all $x \colon S \to E$ and $s \in S$, $x \bullet e_s = x_s$.</li> <li>For every pair of finite sets $S$ and $T$ and all $x \colon S \to E$, $a \colon T \to \mathbb{A} S$ and $b \in \mathbb{A} T$, $x \bullet (a \bullet b) = (x \bullet a) \bullet b$.</li> <li>For every function between finite sets $\phi \colon S \to S'$ and all $x \colon S' \to E$ and $a \in \mathbb{A} S$, $(x \phi) \bullet a = x \bullet (\phi a)$.</li> </ol> <p>In particular, $\mathbb{A} T$ for a fixed $T$ carries a standard structure of an affine space given by the structure maps of $\mathbb{A}$. In this case, the axioms specialize to some standard facts about formal affine combinations and they are in fact the same as axioms saying that $\mathbb{A}$ is a clone in the first place. This way you can recover the full definition of a clone which I neglected to write down above.</p> Answer 2: <p>For $K=\mathbb{R}$, the positive part of your operad (mentioned in Gabriel's comment), and its algebras have been discussed by Tom Leinster and others in connection to entropy. See, for example,</p> <p><a href="https://golem.ph.utexas.edu/category/2011/05/an_operadic_introduction_to_en.html" rel="noreferrer">https://golem.ph.utexas.edu/category/2011/05/an_operadic_introduction_to_en.html</a></p> <p><a href="http://www.maths.ed.ac.uk/~tl/b.pdf" rel="noreferrer">http://www.maths.ed.ac.uk/~tl/b.pdf</a></p> Answer 3: <p>In order to present affine spaces in this fashion you need a variation of operads called <em>clones</em> or <em>cartesian operads</em> (which eventually are equivalent to Lawvere theories). You can find some references on <a href="https://ncatlab.org/nlab/show/clone">nLab</a>.</p> <p>A clone is like a symmetric operad, but with all maps between finite sets acting not just bijections. However, it is more convenient to phrase the definition as follows. A clone is a functor $C \colon \mathsf{FinSet} \to \mathsf{Set}$ equipped with a family of substitution operators, i.e. for each pair of finite sets $S$ and $T$ we specify a function $(C S)^T \times C T \to C S$ and distinguish projection operations $\pi_s \in C S$ for each $s \in S$. These are supposed to satisfy some axioms (see below).</p> <p>I will call the clone for affine spaces $\mathbb{A}$. (And I will refrain from using vocabulary from probability theory, this would go better with convex spaces.) For each $S$, $\mathbb{A} S$ is the set of formal affine combinations of elements of $S$, functoriality of $\mathbb{A}$ is given by grouping coefficients of points together when they are identified by a function between finite sets. Substitution operations evaluate affine combinations of affine combinations. If I have a combination $b \in \mathbb{A} T$ and a family $a \colon T \to \mathbb{A}S$, I will denote the resulting substitution by $a \bullet b \in \mathbb{A}S$. (It is really just a shorthand for $s \mapsto \sum_{t \in T} a_{t, s} b_t$.) Moreover, $\pi_s = e_s$ is the trivial combination with coefficient $1$ at $s$ and $0$ elsewhere.</p> <p>We can now give the following definition. An affine space is a set $E$ equipped with operations $E^S \times \mathbb{A}S \to E$, $(x , a) \mapsto x \bullet a$ subject to the following axioms.</p> <ol> <li>For every finite set $S$ and all $x \colon S \to E$ and $s \in S$, $x \bullet e_s = x_s$.</li> <li>For every pair of finite sets $S$ and $T$ and all $x \colon S \to E$, $a \colon T \to \mathbb{A} S$ and $b \in \mathbb{A} T$, $x \bullet (a \bullet b) = (x \bullet a) \bullet b$.</li> <li>For every function between finite sets $\phi \colon S \to S'$ and all $x \colon S' \to E$ and $a \in \mathbb{A} S$, $(x \phi) \bullet a = x \bullet (\phi a)$.</li> </ol> <p>In particular, $\mathbb{A} T$ for a fixed $T$ carries a standard structure of an affine space given by the structure maps of $\mathbb{A}$. In this case, the axioms specialize to some standard facts about formal affine combinations and they are in fact the same as axioms saying that $\mathbb{A}$ is a clone in the first place. This way you can recover the full definition of a clone which I neglected to write down above.</p> Answer 4: <p>For $K=\mathbb{R}$, the positive part of your operad (mentioned in Gabriel's comment), and its algebras have been discussed by Tom Leinster and others in connection to entropy. See, for example,</p> <p><a href="https://golem.ph.utexas.edu/category/2011/05/an_operadic_introduction_to_en.html" rel="noreferrer">https://golem.ph.utexas.edu/category/2011/05/an_operadic_introduction_to_en.html</a></p> <p><a href="http://www.maths.ed.ac.uk/~tl/b.pdf" rel="noreferrer">http://www.maths.ed.ac.uk/~tl/b.pdf</a></p> Answer 5: <p>In order to present affine spaces in this fashion you need a variation of operads called <em>clones</em> or <em>cartesian operads</em> (which eventually are equivalent to Lawvere theories). You can find some references on <a href="https://ncatlab.org/nlab/show/clone">nLab</a>.</p> <p>A clone is like a symmetric operad, but with all maps between finite sets acting not just bijections. However, it is more convenient to phrase the definition as follows. A clone is a functor $C \colon \mathsf{FinSet} \to \mathsf{Set}$ equipped with a family of substitution operators, i.e. for each pair of finite sets $S$ and $T$ we specify a function $(C S)^T \times C T \to C S$ and distinguish projection operations $\pi_s \in C S$ for each $s \in S$. These are supposed to satisfy some axioms (see below).</p> <p>I will call the clone for affine spaces $\mathbb{A}$. (And I will refrain from using vocabulary from probability theory, this would go better with convex spaces.) For each $S$, $\mathbb{A} S$ is the set of formal affine combinations of elements of $S$, functoriality of $\mathbb{A}$ is given by grouping coefficients of points together when they are identified by a function between finite sets. Substitution operations evaluate affine combinations of affine combinations. If I have a combination $b \in \mathbb{A} T$ and a family $a \colon T \to \mathbb{A}S$, I will denote the resulting substitution by $a \bullet b \in \mathbb{A}S$. (It is really just a shorthand for $s \mapsto \sum_{t \in T} a_{t, s} b_t$.) Moreover, $\pi_s = e_s$ is the trivial combination with coefficient $1$ at $s$ and $0$ elsewhere.</p> <p>We can now give the following definition. An affine space is a set $E$ equipped with operations $E^S \times \mathbb{A}S \to E$, $(x , a) \mapsto x \bullet a$ subject to the following axioms.</p> <ol> <li>For every finite set $S$ and all $x \colon S \to E$ and $s \in S$, $x \bullet e_s = x_s$.</li> <li>For every pair of finite sets $S$ and $T$ and all $x \colon S \to E$, $a \colon T \to \mathbb{A} S$ and $b \in \mathbb{A} T$, $x \bullet (a \bullet b) = (x \bullet a) \bullet b$.</li> <li>For every function between finite sets $\phi \colon S \to S'$ and all $x \colon S' \to E$ and $a \in \mathbb{A} S$, $(x \phi) \bullet a = x \bullet (\phi a)$.</li> </ol> <p>In particular, $\mathbb{A} T$ for a fixed $T$ carries a standard structure of an affine space given by the structure maps of $\mathbb{A}$. In this case, the axioms specialize to some standard facts about formal affine combinations and they are in fact the same as axioms saying that $\mathbb{A}$ is a clone in the first place. This way you can recover the full definition of a clone which I neglected to write down above.</p> Answer 6: <p>For $K=\mathbb{R}$, the positive part of your operad (mentioned in Gabriel's comment), and its algebras have been discussed by Tom Leinster and others in connection to entropy. See, for example,</p> <p><a href="https://golem.ph.utexas.edu/category/2011/05/an_operadic_introduction_to_en.html" rel="noreferrer">https://golem.ph.utexas.edu/category/2011/05/an_operadic_introduction_to_en.html</a></p> <p><a href="http://www.maths.ed.ac.uk/~tl/b.pdf" rel="noreferrer">http://www.maths.ed.ac.uk/~tl/b.pdf</a></p>
Choosing subsets of $\mathbb R$ of cardinality $\frak c$, who wins?
258,597
9
https://mathoverflow.net/questions/258597/choosing-subsets-of-mathbb-r-of-cardinality-frak-c-who-wins
<p>Consider the following infinite game: two players, I and II, are alternating and choosing a descending sequence of subsets of $\mathbb R$ of cardinality $\frak c$, so I chooses a set $A_1\subseteq\mathbb R$, II chooses a set $A_2\subseteq A_1$, I chooses $A_3\subseteq A_2$ etc., all having the size continuum. Let $A=\cap_{i=1}^\infty A_i$. I wins if $A$ is nonempty, otherwise II wins.</p> <blockquote> <p>Who has a winning strategy in this game?</p> </blockquote> <p>This game was posed to me as a puzzle from a person I don't have any means of contact with anymore (so I can't ask them about this). However, I don't know whether he himself knew the answer or not (as far as I know it might be independent of ZFC or unknown) which is why I'm posting this on MO and not Math.SE (the same person has posed as a puzzle a combinatorial problem which I happened to know is open).</p> <p>All I know is that it's consistent with ZF that II has the winning strategy: if $\mathbb R$ had countable cofinality, i.e. was a countable union of sets $S_i,i\in\mathbb N$ of cardinality $&lt;\frak c$, then II could play on his turns $A_{2k}=A_{2k-1}\setminus(S_1\cup\dots\cup S_n)$, which would have size $\frak c$. This doesn't go through in ZFC, since there $\frak c$ has uncountable cofinality.</p> <p>One can also consider the problem with $\mathbb R$ replaced by some other set, for example an ordinal. For some sufficiently well-behaved sets (not Dedekind-finite, for example) the above reasoning shows that if a cardinal has countable cofinality, I has a winning strategy. I am especially interested in the case of $\omega_1$, which is equivalent to the problem at hand if we assume CH.</p> <p>I am interested in this problem both in context of ZF and ZFC, and I am willing to assume CH to resolve the original problem if it is of help there.</p>
[ { "answer_id": 258598, "score": 13, "text": "<p>In ZFC, the player aiming for the empty set has a winning strategy in the game played on any infinite set, including the reals. Using the axiom of choice, we can well-order the set and thereby pretend that we are playing sets of ordinals. Now, whenever it is the empty-set player's turn, she should look at the order type of the ordinals remaining, and play so as to omit the first element of every successive block of $\\omega$ many elements in the enumeration of the set. In this way, no element can survive into the final intersection. If an element is the $(k+1)^{th}$ in it's $\\omega$-block, then it is the $(\\omega\\cdot\\gamma+k+1)^{th}$ element of $A_n$ for some ordinal $\\gamma$, and after the empty-set player plays, it will become the $(\\omega\\cdot\\gamma+k)^{th}$ element of the new set, strictly smaller. When the other player plays, if it was the $\\alpha^{th}$ element, it becomes the $\\beta^{th}$ element for some $\\beta\\leq\\alpha$. So as play progresses, the index of the element in the enumeration will steadily descend. Since it cannot descend forever, the element must eventually fall out of the set and so the final intersection is empty, as desired. </p>\n\n<p>In ZF, <a href=\"https://mathoverflow.net/a/137367/1946\">the argument I gave at the other question</a> shows that the non-empty player cannot have a winning strategy in the game on the reals. </p>\n" }, { "answer_id": 258649, "score": 8, "text": "<p>This problem (with inconsequential differences in the rules of play) was posed by S. Banach as Problem 67 in <a href=\"https://en.wikipedia.org/wiki/Scottish_Book\" rel=\"nofollow noreferrer\"><em>The Scottish Book</em></a>, which is available <a href=\"http://kielich.amu.edu.pl/Stefan_Banach/pdf/ks-szkocka/ks-szkocka3ang.pdf\" rel=\"nofollow noreferrer\">here</a>. The following text is copied from <a href=\"https://books.google.com/books?id=gaqEAAAAIAAJ\" rel=\"nofollow noreferrer\">R. Daniel Mauldin's edition</a>:</p>\n\n<blockquote>\n <p>(A MODIFICATION OF MAZUR'S game, [see Problem 43]).<br>\n <br>We call a half of the set $E$ [in symbols, $(1/2)E$] an arbitrary subset $H\\subset E$ such that the sets $E,H,E-H$ are of equal power.<br>\n <br>(1) Two players $A$ and $B$ give in turn sets $E_i, i=1,2,\\dots$ ad inf. so that $E_i=(1/2)E_{i-1}\\ i=1,2,\\dots$ where $E_0$ is a given abstract set. Player $A$ wins if the product $E_1E_2\\dots E_iE_{i+1}\\dots$ is vacuous.<br>\n <br>(2) The game, similar to the one above, with the assumption that $E_i=1/2[E_0-E_1-\\dots-E_{i-1}]\\ i=2,3,\\dots$ ad inf., and $E_1=(1/2)E_0.$ Player $A$ wins if $E_1+E_2+\\dots=E_0.$<br>\n <br>Is there a method of win for player $A$? If $E_0$ is of power cofinal with $\\aleph_0,$ then player $A$ has a method of win. Is it only in this case? In particular, solve the problem if $E_0$ is the set of real numbers.<br>\n <br><em>Addendum</em>. There exists a method of play which will guarantee that the product of the sets is not vacuous. The solution was given by J. Schreier.</p>\n</blockquote>\n\n<p>The problem is dated August 1, 1935; the addendum is dated August 24, 1935; Schreier's solution (using the well-ordering theorem, see <a href=\"https://mathoverflow.net/questions/258597/choosing-subsets-of-mathbb-r-of-cardinality-frak-c-who-wins/258598#258598\">Joel David Hamkins' answer</a>) was published in the paper</p>\n\n<blockquote>\n <p>J. Schreier, <a href=\"http://matwbn.icm.edu.pl/ksiazki/sm/sm7/sm7115.pdf\" rel=\"nofollow noreferrer\">Eine Eigenschaft abstrakter Mengen</a>, <em>Studia Math.</em> <strong>7</strong> (1938), 155–156, <a href=\"https://eudml.org/doc/218619\" rel=\"nofollow noreferrer\">eudml</a>.</p>\n</blockquote>\n" }, { "answer_id": 258598, "score": 13, "text": "<p>In ZFC, the player aiming for the empty set has a winning strategy in the game played on any infinite set, including the reals. Using the axiom of choice, we can well-order the set and thereby pretend that we are playing sets of ordinals. Now, whenever it is the empty-set player's turn, she should look at the order type of the ordinals remaining, and play so as to omit the first element of every successive block of $\\omega$ many elements in the enumeration of the set. In this way, no element can survive into the final intersection. If an element is the $(k+1)^{th}$ in it's $\\omega$-block, then it is the $(\\omega\\cdot\\gamma+k+1)^{th}$ element of $A_n$ for some ordinal $\\gamma$, and after the empty-set player plays, it will become the $(\\omega\\cdot\\gamma+k)^{th}$ element of the new set, strictly smaller. When the other player plays, if it was the $\\alpha^{th}$ element, it becomes the $\\beta^{th}$ element for some $\\beta\\leq\\alpha$. So as play progresses, the index of the element in the enumeration will steadily descend. Since it cannot descend forever, the element must eventually fall out of the set and so the final intersection is empty, as desired. </p>\n\n<p>In ZF, <a href=\"https://mathoverflow.net/a/137367/1946\">the argument I gave at the other question</a> shows that the non-empty player cannot have a winning strategy in the game on the reals. </p>\n" }, { "answer_id": 258649, "score": 8, "text": "<p>This problem (with inconsequential differences in the rules of play) was posed by S. Banach as Problem 67 in <a href=\"https://en.wikipedia.org/wiki/Scottish_Book\" rel=\"nofollow noreferrer\"><em>The Scottish Book</em></a>, which is available <a href=\"http://kielich.amu.edu.pl/Stefan_Banach/pdf/ks-szkocka/ks-szkocka3ang.pdf\" rel=\"nofollow noreferrer\">here</a>. The following text is copied from <a href=\"https://books.google.com/books?id=gaqEAAAAIAAJ\" rel=\"nofollow noreferrer\">R. Daniel Mauldin's edition</a>:</p>\n\n<blockquote>\n <p>(A MODIFICATION OF MAZUR'S game, [see Problem 43]).<br>\n <br>We call a half of the set $E$ [in symbols, $(1/2)E$] an arbitrary subset $H\\subset E$ such that the sets $E,H,E-H$ are of equal power.<br>\n <br>(1) Two players $A$ and $B$ give in turn sets $E_i, i=1,2,\\dots$ ad inf. so that $E_i=(1/2)E_{i-1}\\ i=1,2,\\dots$ where $E_0$ is a given abstract set. Player $A$ wins if the product $E_1E_2\\dots E_iE_{i+1}\\dots$ is vacuous.<br>\n <br>(2) The game, similar to the one above, with the assumption that $E_i=1/2[E_0-E_1-\\dots-E_{i-1}]\\ i=2,3,\\dots$ ad inf., and $E_1=(1/2)E_0.$ Player $A$ wins if $E_1+E_2+\\dots=E_0.$<br>\n <br>Is there a method of win for player $A$? If $E_0$ is of power cofinal with $\\aleph_0,$ then player $A$ has a method of win. Is it only in this case? In particular, solve the problem if $E_0$ is the set of real numbers.<br>\n <br><em>Addendum</em>. There exists a method of play which will guarantee that the product of the sets is not vacuous. The solution was given by J. Schreier.</p>\n</blockquote>\n\n<p>The problem is dated August 1, 1935; the addendum is dated August 24, 1935; Schreier's solution (using the well-ordering theorem, see <a href=\"https://mathoverflow.net/questions/258597/choosing-subsets-of-mathbb-r-of-cardinality-frak-c-who-wins/258598#258598\">Joel David Hamkins' answer</a>) was published in the paper</p>\n\n<blockquote>\n <p>J. Schreier, <a href=\"http://matwbn.icm.edu.pl/ksiazki/sm/sm7/sm7115.pdf\" rel=\"nofollow noreferrer\">Eine Eigenschaft abstrakter Mengen</a>, <em>Studia Math.</em> <strong>7</strong> (1938), 155–156, <a href=\"https://eudml.org/doc/218619\" rel=\"nofollow noreferrer\">eudml</a>.</p>\n</blockquote>\n" } ]
Question: Choosing subsets of $\mathbb R$ of cardinality $\frak c$, who wins? Answer 1: <p>In ZFC, the player aiming for the empty set has a winning strategy in the game played on any infinite set, including the reals. Using the axiom of choice, we can well-order the set and thereby pretend that we are playing sets of ordinals. Now, whenever it is the empty-set player's turn, she should look at the order type of the ordinals remaining, and play so as to omit the first element of every successive block of $\omega$ many elements in the enumeration of the set. In this way, no element can survive into the final intersection. If an element is the $(k+1)^{th}$ in it's $\omega$-block, then it is the $(\omega\cdot\gamma+k+1)^{th}$ element of $A_n$ for some ordinal $\gamma$, and after the empty-set player plays, it will become the $(\omega\cdot\gamma+k)^{th}$ element of the new set, strictly smaller. When the other player plays, if it was the $\alpha^{th}$ element, it becomes the $\beta^{th}$ element for some $\beta\leq\alpha$. So as play progresses, the index of the element in the enumeration will steadily descend. Since it cannot descend forever, the element must eventually fall out of the set and so the final intersection is empty, as desired. </p> <p>In ZF, <a href="https://mathoverflow.net/a/137367/1946">the argument I gave at the other question</a> shows that the non-empty player cannot have a winning strategy in the game on the reals. </p> Answer 2: <p>This problem (with inconsequential differences in the rules of play) was posed by S. Banach as Problem 67 in <a href="https://en.wikipedia.org/wiki/Scottish_Book" rel="nofollow noreferrer"><em>The Scottish Book</em></a>, which is available <a href="http://kielich.amu.edu.pl/Stefan_Banach/pdf/ks-szkocka/ks-szkocka3ang.pdf" rel="nofollow noreferrer">here</a>. The following text is copied from <a href="https://books.google.com/books?id=gaqEAAAAIAAJ" rel="nofollow noreferrer">R. Daniel Mauldin's edition</a>:</p> <blockquote> <p>(A MODIFICATION OF MAZUR'S game, [see Problem 43]).<br> <br>We call a half of the set $E$ [in symbols, $(1/2)E$] an arbitrary subset $H\subset E$ such that the sets $E,H,E-H$ are of equal power.<br> <br>(1) Two players $A$ and $B$ give in turn sets $E_i, i=1,2,\dots$ ad inf. so that $E_i=(1/2)E_{i-1}\ i=1,2,\dots$ where $E_0$ is a given abstract set. Player $A$ wins if the product $E_1E_2\dots E_iE_{i+1}\dots$ is vacuous.<br> <br>(2) The game, similar to the one above, with the assumption that $E_i=1/2[E_0-E_1-\dots-E_{i-1}]\ i=2,3,\dots$ ad inf., and $E_1=(1/2)E_0.$ Player $A$ wins if $E_1+E_2+\dots=E_0.$<br> <br>Is there a method of win for player $A$? If $E_0$ is of power cofinal with $\aleph_0,$ then player $A$ has a method of win. Is it only in this case? In particular, solve the problem if $E_0$ is the set of real numbers.<br> <br><em>Addendum</em>. There exists a method of play which will guarantee that the product of the sets is not vacuous. The solution was given by J. Schreier.</p> </blockquote> <p>The problem is dated August 1, 1935; the addendum is dated August 24, 1935; Schreier's solution (using the well-ordering theorem, see <a href="https://mathoverflow.net/questions/258597/choosing-subsets-of-mathbb-r-of-cardinality-frak-c-who-wins/258598#258598">Joel David Hamkins' answer</a>) was published in the paper</p> <blockquote> <p>J. Schreier, <a href="http://matwbn.icm.edu.pl/ksiazki/sm/sm7/sm7115.pdf" rel="nofollow noreferrer">Eine Eigenschaft abstrakter Mengen</a>, <em>Studia Math.</em> <strong>7</strong> (1938), 155–156, <a href="https://eudml.org/doc/218619" rel="nofollow noreferrer">eudml</a>.</p> </blockquote> Answer 3: <p>In ZFC, the player aiming for the empty set has a winning strategy in the game played on any infinite set, including the reals. Using the axiom of choice, we can well-order the set and thereby pretend that we are playing sets of ordinals. Now, whenever it is the empty-set player's turn, she should look at the order type of the ordinals remaining, and play so as to omit the first element of every successive block of $\omega$ many elements in the enumeration of the set. In this way, no element can survive into the final intersection. If an element is the $(k+1)^{th}$ in it's $\omega$-block, then it is the $(\omega\cdot\gamma+k+1)^{th}$ element of $A_n$ for some ordinal $\gamma$, and after the empty-set player plays, it will become the $(\omega\cdot\gamma+k)^{th}$ element of the new set, strictly smaller. When the other player plays, if it was the $\alpha^{th}$ element, it becomes the $\beta^{th}$ element for some $\beta\leq\alpha$. So as play progresses, the index of the element in the enumeration will steadily descend. Since it cannot descend forever, the element must eventually fall out of the set and so the final intersection is empty, as desired. </p> <p>In ZF, <a href="https://mathoverflow.net/a/137367/1946">the argument I gave at the other question</a> shows that the non-empty player cannot have a winning strategy in the game on the reals. </p> Answer 4: <p>This problem (with inconsequential differences in the rules of play) was posed by S. Banach as Problem 67 in <a href="https://en.wikipedia.org/wiki/Scottish_Book" rel="nofollow noreferrer"><em>The Scottish Book</em></a>, which is available <a href="http://kielich.amu.edu.pl/Stefan_Banach/pdf/ks-szkocka/ks-szkocka3ang.pdf" rel="nofollow noreferrer">here</a>. The following text is copied from <a href="https://books.google.com/books?id=gaqEAAAAIAAJ" rel="nofollow noreferrer">R. Daniel Mauldin's edition</a>:</p> <blockquote> <p>(A MODIFICATION OF MAZUR'S game, [see Problem 43]).<br> <br>We call a half of the set $E$ [in symbols, $(1/2)E$] an arbitrary subset $H\subset E$ such that the sets $E,H,E-H$ are of equal power.<br> <br>(1) Two players $A$ and $B$ give in turn sets $E_i, i=1,2,\dots$ ad inf. so that $E_i=(1/2)E_{i-1}\ i=1,2,\dots$ where $E_0$ is a given abstract set. Player $A$ wins if the product $E_1E_2\dots E_iE_{i+1}\dots$ is vacuous.<br> <br>(2) The game, similar to the one above, with the assumption that $E_i=1/2[E_0-E_1-\dots-E_{i-1}]\ i=2,3,\dots$ ad inf., and $E_1=(1/2)E_0.$ Player $A$ wins if $E_1+E_2+\dots=E_0.$<br> <br>Is there a method of win for player $A$? If $E_0$ is of power cofinal with $\aleph_0,$ then player $A$ has a method of win. Is it only in this case? In particular, solve the problem if $E_0$ is the set of real numbers.<br> <br><em>Addendum</em>. There exists a method of play which will guarantee that the product of the sets is not vacuous. The solution was given by J. Schreier.</p> </blockquote> <p>The problem is dated August 1, 1935; the addendum is dated August 24, 1935; Schreier's solution (using the well-ordering theorem, see <a href="https://mathoverflow.net/questions/258597/choosing-subsets-of-mathbb-r-of-cardinality-frak-c-who-wins/258598#258598">Joel David Hamkins' answer</a>) was published in the paper</p> <blockquote> <p>J. Schreier, <a href="http://matwbn.icm.edu.pl/ksiazki/sm/sm7/sm7115.pdf" rel="nofollow noreferrer">Eine Eigenschaft abstrakter Mengen</a>, <em>Studia Math.</em> <strong>7</strong> (1938), 155–156, <a href="https://eudml.org/doc/218619" rel="nofollow noreferrer">eudml</a>.</p> </blockquote>
Is there a useful limit or co-limit of a diagram that has only a single object?
258,855
9
https://mathoverflow.net/questions/258855/is-there-a-useful-limit-or-co-limit-of-a-diagram-that-has-only-a-single-object
<p>I'm starting to study category theory kind of informally and everytime I read about the definitions of limits and co-limits, the first three examples are always the same: </p> <ul> <li>terminal/initial objects,</li> <li>products/coproducts,</li> <li>and pullbacks/pushouts. </li> </ul> <p>It's always explained how terminal (initial) objects are the limit (colimit) of a an empty diagram and products (coproducts) are the limit (colimit) of a diagram with only two objects with no additional structure.</p> <p>I always get that feeling of "what about the diagram with a single object?", but since that's never mentioned, I figured it must be something trivial or not useful. In a Poset it feels it would be some kind of lowest upper bound/greatest lower bound of a single object or something like that, but that looks like it would be trivially equal to the object itself. <strong>So, first question: is this limit interesting or is it trivial or maybe not used anywhere?</strong></p> <p>A related question: Recently I started reading studying the book Mathematical Physics by Robert Geroch (*) and when he defines a free group it feels a lot like what the limit of a diagram with a single object would be: "a free group on the set $S$ is a group $G$ together with a mapping $\alpha$ from $S$ to $G$ such that for any other group $G'$ with a mapping $\alpha'$ from $S$ to $G'$, there exists a unique homomorphism $\mu: G \to G'$ such that the diagram commutes", meaning that $\alpha' = \mu \cdot \alpha$. But of course that's not a limit because G and S belong to different categories and $\alpha$ and $\alpha'$ are not morphisms. The books defines other kinds of free things after that in a similar way -- free vector spaces, free topological spaces, etc. <strong>So, the second question is: are those free constructions related to limits? Is there some category where those constructions are actual limits of a diagram with a single object?</strong></p> <p>(*) I'm a physicist and my mathematical training was always focused on mathematical methods for physics and very little in formal, rigorous mathematics - since this book doesn't assume a lot of education in things like topology, algebraic geometric, etc, it feels more adequate for my background. Would love other suggestions.</p>
[ { "answer_id": 258857, "score": 23, "text": "<p>One example: Let $X$ be a set with a group $G$ acting on it. Consider the diagram, in the category of sets, with $X$ as its only object, but with all the elements of $G$ (considered as permutations of $X$) as morphisms. The limit of this diagram is the subset of $X$ consisting of the points fixed by the group action, and the colimit is the set of orbits of the action.</p>\n" }, { "answer_id": 258925, "score": 2, "text": "<p>Let $|G|$ be set of elements of a group $G$. Similarly, if $f : G \\to H$ is a group homomorphism, let $|f|$ be the underlying mapping on elements. A diagrammatic translation of that definition of free group is as follows.</p>\n\n<p>Consider the category whose objects are:</p>\n\n<ul>\n<li>Pairs $(G',\\alpha')$ consisting of a group $G'$ together with a function of sets $\\alpha' : S \\to |G'|$</li>\n<li>Arrows $(G',\\alpha') \\to (G'', \\alpha'')$ are group homomorphisms $\\beta : G' \\to G''$ satisfying $|\\beta| \\circ \\alpha' = \\alpha''$</li>\n</ul>\n\n<p>It turns out this category has an initial object (which can be seen either as the colimit of an empty diagram or the limit of the entire category), and the pair $(G,\\alpha)$ is taken to be an initial object.</p>\n\n<p>So, I think your intuition was mashing together two separate facts:</p>\n\n<ul>\n<li>The free group is indeed the colimit of a fairly small diagram &mdash; the empty one!</li>\n<li>All of the objects in this category involve maps from the single set $S$ </li>\n</ul>\n\n<hr>\n\n<p>Incidentally, this sort of pattern comes up a lot &mdash; constructing a category whose objects are collections of interesting data, and then doing things in that category.</p>\n\n<p>The particular category above is an example of a comma category from the constant functor $* \\to \\mathbf{Set}$ whose image is $S$ to the forgetful functor $\\mathbf{Grp} \\to \\mathbf{Set}$. In fact, one of the things you can do with comma categories is precisely to make sense of arrows between objects of different categories.</p>\n\n<p>The construction above is also an example of an adjunction; the \"free group\" functor $\\mathbf{Set} \\to \\mathbf{Grp}$ is the left adjoint to the \"underlying set of elements\" functor $\\mathbf{Grp} \\to \\mathbf{Set}$.</p>\n" }, { "answer_id": 258857, "score": 23, "text": "<p>One example: Let $X$ be a set with a group $G$ acting on it. Consider the diagram, in the category of sets, with $X$ as its only object, but with all the elements of $G$ (considered as permutations of $X$) as morphisms. The limit of this diagram is the subset of $X$ consisting of the points fixed by the group action, and the colimit is the set of orbits of the action.</p>\n" }, { "answer_id": 258925, "score": 2, "text": "<p>Let $|G|$ be set of elements of a group $G$. Similarly, if $f : G \\to H$ is a group homomorphism, let $|f|$ be the underlying mapping on elements. A diagrammatic translation of that definition of free group is as follows.</p>\n\n<p>Consider the category whose objects are:</p>\n\n<ul>\n<li>Pairs $(G',\\alpha')$ consisting of a group $G'$ together with a function of sets $\\alpha' : S \\to |G'|$</li>\n<li>Arrows $(G',\\alpha') \\to (G'', \\alpha'')$ are group homomorphisms $\\beta : G' \\to G''$ satisfying $|\\beta| \\circ \\alpha' = \\alpha''$</li>\n</ul>\n\n<p>It turns out this category has an initial object (which can be seen either as the colimit of an empty diagram or the limit of the entire category), and the pair $(G,\\alpha)$ is taken to be an initial object.</p>\n\n<p>So, I think your intuition was mashing together two separate facts:</p>\n\n<ul>\n<li>The free group is indeed the colimit of a fairly small diagram &mdash; the empty one!</li>\n<li>All of the objects in this category involve maps from the single set $S$ </li>\n</ul>\n\n<hr>\n\n<p>Incidentally, this sort of pattern comes up a lot &mdash; constructing a category whose objects are collections of interesting data, and then doing things in that category.</p>\n\n<p>The particular category above is an example of a comma category from the constant functor $* \\to \\mathbf{Set}$ whose image is $S$ to the forgetful functor $\\mathbf{Grp} \\to \\mathbf{Set}$. In fact, one of the things you can do with comma categories is precisely to make sense of arrows between objects of different categories.</p>\n\n<p>The construction above is also an example of an adjunction; the \"free group\" functor $\\mathbf{Set} \\to \\mathbf{Grp}$ is the left adjoint to the \"underlying set of elements\" functor $\\mathbf{Grp} \\to \\mathbf{Set}$.</p>\n" } ]
Question: Is there a useful limit or co-limit of a diagram that has only a single object? Answer 1: <p>One example: Let $X$ be a set with a group $G$ acting on it. Consider the diagram, in the category of sets, with $X$ as its only object, but with all the elements of $G$ (considered as permutations of $X$) as morphisms. The limit of this diagram is the subset of $X$ consisting of the points fixed by the group action, and the colimit is the set of orbits of the action.</p> Answer 2: <p>Let $|G|$ be set of elements of a group $G$. Similarly, if $f : G \to H$ is a group homomorphism, let $|f|$ be the underlying mapping on elements. A diagrammatic translation of that definition of free group is as follows.</p> <p>Consider the category whose objects are:</p> <ul> <li>Pairs $(G',\alpha')$ consisting of a group $G'$ together with a function of sets $\alpha' : S \to |G'|$</li> <li>Arrows $(G',\alpha') \to (G'', \alpha'')$ are group homomorphisms $\beta : G' \to G''$ satisfying $|\beta| \circ \alpha' = \alpha''$</li> </ul> <p>It turns out this category has an initial object (which can be seen either as the colimit of an empty diagram or the limit of the entire category), and the pair $(G,\alpha)$ is taken to be an initial object.</p> <p>So, I think your intuition was mashing together two separate facts:</p> <ul> <li>The free group is indeed the colimit of a fairly small diagram &mdash; the empty one!</li> <li>All of the objects in this category involve maps from the single set $S$ </li> </ul> <hr> <p>Incidentally, this sort of pattern comes up a lot &mdash; constructing a category whose objects are collections of interesting data, and then doing things in that category.</p> <p>The particular category above is an example of a comma category from the constant functor $* \to \mathbf{Set}$ whose image is $S$ to the forgetful functor $\mathbf{Grp} \to \mathbf{Set}$. In fact, one of the things you can do with comma categories is precisely to make sense of arrows between objects of different categories.</p> <p>The construction above is also an example of an adjunction; the "free group" functor $\mathbf{Set} \to \mathbf{Grp}$ is the left adjoint to the "underlying set of elements" functor $\mathbf{Grp} \to \mathbf{Set}$.</p> Answer 3: <p>One example: Let $X$ be a set with a group $G$ acting on it. Consider the diagram, in the category of sets, with $X$ as its only object, but with all the elements of $G$ (considered as permutations of $X$) as morphisms. The limit of this diagram is the subset of $X$ consisting of the points fixed by the group action, and the colimit is the set of orbits of the action.</p> Answer 4: <p>Let $|G|$ be set of elements of a group $G$. Similarly, if $f : G \to H$ is a group homomorphism, let $|f|$ be the underlying mapping on elements. A diagrammatic translation of that definition of free group is as follows.</p> <p>Consider the category whose objects are:</p> <ul> <li>Pairs $(G',\alpha')$ consisting of a group $G'$ together with a function of sets $\alpha' : S \to |G'|$</li> <li>Arrows $(G',\alpha') \to (G'', \alpha'')$ are group homomorphisms $\beta : G' \to G''$ satisfying $|\beta| \circ \alpha' = \alpha''$</li> </ul> <p>It turns out this category has an initial object (which can be seen either as the colimit of an empty diagram or the limit of the entire category), and the pair $(G,\alpha)$ is taken to be an initial object.</p> <p>So, I think your intuition was mashing together two separate facts:</p> <ul> <li>The free group is indeed the colimit of a fairly small diagram &mdash; the empty one!</li> <li>All of the objects in this category involve maps from the single set $S$ </li> </ul> <hr> <p>Incidentally, this sort of pattern comes up a lot &mdash; constructing a category whose objects are collections of interesting data, and then doing things in that category.</p> <p>The particular category above is an example of a comma category from the constant functor $* \to \mathbf{Set}$ whose image is $S$ to the forgetful functor $\mathbf{Grp} \to \mathbf{Set}$. In fact, one of the things you can do with comma categories is precisely to make sense of arrows between objects of different categories.</p> <p>The construction above is also an example of an adjunction; the "free group" functor $\mathbf{Set} \to \mathbf{Grp}$ is the left adjoint to the "underlying set of elements" functor $\mathbf{Grp} \to \mathbf{Set}$.</p>
Efficient method to write number as a sum of four squares?
259,152
9
https://mathoverflow.net/questions/259152/efficient-method-to-write-number-as-a-sum-of-four-squares
<p><a href="https://en.wikipedia.org/wiki/Lagrange&#39;s_four-square_theorem#Algorithms" rel="noreferrer">Wikipedia</a> states that there randomized polynomial-time algorithms for writing $n$ as a sum of four squares</p> <p>$n=x_{1}^{2}+x_{2}^{2}+x_{3}^{2}+x_{4}^{2}$ </p> <p>in expected running time $\mathrm {O} (\log^{2} n).$</p> <p>My question is can someone give the efficient algorithm( $\mathrm {O} (\log^{2} n)$ ) to represent $n$ as sum of four squares.</p>
[ { "answer_id": 270797, "score": 3, "text": "<p>One of their methods for $n=4k+2$ is as follows:</p>\n\n<p>Randomly select an even number $a$ and an odd number $b$ such that $a^2+b^2 &lt; n$. Then, we hope $p=n-a^2-b^2$ is a prime (you can show there's about a $1/(A \\log n \\log \\log n)$ chance of $p$ being prime ); $p$ is of the form $4r+1$, so if $p$ is prime there's a solution to $c^2+d^2=p$.</p>\n\n<p>To find that, we try to solve $m^2+1 \\equiv 0 \\pmod{p}$; I'm actually going to describe a slightly different method. Select $x$ at random from $1$ to $p-1$; then, $x^{2r}=\\pm 1$ depending on whether $x$ is a quadratic residue so calculate $x^r$ by repeated squaring, and with a $1/2$ chance if $p$ is prime (a smaller one if $p$ is composite) you'll find a valid $m$.</p>\n\n<p>Given such an $m$, $m+i$ is a Gaussian integer with norm divisible by $p$ but smaller than $p^2$; use the Euclidean algorithm on the Gaussian integers with $p$ and get $c+di$ with norm $p$, and $a^2+b^2+c^2+d^2=n$.</p>\n\n<p>For an odd number $n$, solve for $a^2+b^2+c^2+d^2=2n$; note that by mod $4$ considerations exactly two of $a,b,c,d$ must be odd, and without loss of generality assume $a,b$ are odd and $c,d$ are even. Then,$(\\frac{1}{2}(a+b))^2+(\\frac{1}{2}(a-b))^2+(\\frac{1}{2}(c+d))^2+(\\frac{1}{2}(c-d))^2=n.$</p>\n\n<p>For $n$ a multiple of $4$, solve for $n/4$ recursively, and multiply all values by $4$.</p>\n" }, { "answer_id": 259158, "score": 0, "text": "<p>This seems to be addressed in <a href=\"https://pdfs.semanticscholar.org/9bd7/f2b063bfd2ec831be9c58750287d8a77b186.pdf\" rel=\"nofollow noreferrer\">this paper by Bumby.</a></p>\n" }, { "answer_id": 259203, "score": 0, "text": "<p>I am going to provide you with 4 links and I hope it will help you.</p>\n\n<p>this one is an online calculator based on an algorithm from math overflow.</p>\n\n<p><a href=\"http://www.mathcelebrity.com/foursquare.php?num=+178&amp;pl=Show+Lagrange+Four-Square+Notation\" rel=\"nofollow noreferrer\">http://www.mathcelebrity.com/foursquare.php?num=+178&amp;pl=Show+Lagrange+Four-Square+Notation</a> </p>\n\n<p>the math overflow algorithm can be found here.</p>\n\n<p><a href=\"https://stackoverflow.com/questions/11732555/how-to-find-all-possible-values-of-four-variables-when-squared-sum-to-n\">https://stackoverflow.com/questions/11732555/how-to-find-all-possible-values-of-four-variables-when-squared-sum-to-n</a> </p>\n\n<p>and then there is also this algorithm from cs stackexchange </p>\n\n<p><a href=\"https://cs.stackexchange.com/questions/2988/how-fast-can-we-find-all-four-square-combinations-that-sum-to-n\">https://cs.stackexchange.com/questions/2988/how-fast-can-we-find-all-four-square-combinations-that-sum-to-n</a></p>\n\n<p>and if you want to do a bit of work, then there is this algorithm that can find the sum of 2 squares for a given integer. So you would have to divide your number in two ( not necessarily equal part ) and then add up the results to get a 4 square representation. </p>\n\n<p><a href=\"https://math.stackexchange.com/questions/1972771/is-this-the-general-solution-of-finding-the-two-original-squares-that-add-up-to/1977767?noredirect=1#comment4062390_1977767\">https://math.stackexchange.com/questions/1972771/is-this-the-general-solution-of-finding-the-two-original-squares-that-add-up-to/1977767?noredirect=1#comment4062390_1977767</a></p>\n" }, { "answer_id": 270797, "score": 3, "text": "<p>One of their methods for $n=4k+2$ is as follows:</p>\n\n<p>Randomly select an even number $a$ and an odd number $b$ such that $a^2+b^2 &lt; n$. Then, we hope $p=n-a^2-b^2$ is a prime (you can show there's about a $1/(A \\log n \\log \\log n)$ chance of $p$ being prime ); $p$ is of the form $4r+1$, so if $p$ is prime there's a solution to $c^2+d^2=p$.</p>\n\n<p>To find that, we try to solve $m^2+1 \\equiv 0 \\pmod{p}$; I'm actually going to describe a slightly different method. Select $x$ at random from $1$ to $p-1$; then, $x^{2r}=\\pm 1$ depending on whether $x$ is a quadratic residue so calculate $x^r$ by repeated squaring, and with a $1/2$ chance if $p$ is prime (a smaller one if $p$ is composite) you'll find a valid $m$.</p>\n\n<p>Given such an $m$, $m+i$ is a Gaussian integer with norm divisible by $p$ but smaller than $p^2$; use the Euclidean algorithm on the Gaussian integers with $p$ and get $c+di$ with norm $p$, and $a^2+b^2+c^2+d^2=n$.</p>\n\n<p>For an odd number $n$, solve for $a^2+b^2+c^2+d^2=2n$; note that by mod $4$ considerations exactly two of $a,b,c,d$ must be odd, and without loss of generality assume $a,b$ are odd and $c,d$ are even. Then,$(\\frac{1}{2}(a+b))^2+(\\frac{1}{2}(a-b))^2+(\\frac{1}{2}(c+d))^2+(\\frac{1}{2}(c-d))^2=n.$</p>\n\n<p>For $n$ a multiple of $4$, solve for $n/4$ recursively, and multiply all values by $4$.</p>\n" }, { "answer_id": 259158, "score": 0, "text": "<p>This seems to be addressed in <a href=\"https://pdfs.semanticscholar.org/9bd7/f2b063bfd2ec831be9c58750287d8a77b186.pdf\" rel=\"nofollow noreferrer\">this paper by Bumby.</a></p>\n" }, { "answer_id": 259203, "score": 0, "text": "<p>I am going to provide you with 4 links and I hope it will help you.</p>\n\n<p>this one is an online calculator based on an algorithm from math overflow.</p>\n\n<p><a href=\"http://www.mathcelebrity.com/foursquare.php?num=+178&amp;pl=Show+Lagrange+Four-Square+Notation\" rel=\"nofollow noreferrer\">http://www.mathcelebrity.com/foursquare.php?num=+178&amp;pl=Show+Lagrange+Four-Square+Notation</a> </p>\n\n<p>the math overflow algorithm can be found here.</p>\n\n<p><a href=\"https://stackoverflow.com/questions/11732555/how-to-find-all-possible-values-of-four-variables-when-squared-sum-to-n\">https://stackoverflow.com/questions/11732555/how-to-find-all-possible-values-of-four-variables-when-squared-sum-to-n</a> </p>\n\n<p>and then there is also this algorithm from cs stackexchange </p>\n\n<p><a href=\"https://cs.stackexchange.com/questions/2988/how-fast-can-we-find-all-four-square-combinations-that-sum-to-n\">https://cs.stackexchange.com/questions/2988/how-fast-can-we-find-all-four-square-combinations-that-sum-to-n</a></p>\n\n<p>and if you want to do a bit of work, then there is this algorithm that can find the sum of 2 squares for a given integer. So you would have to divide your number in two ( not necessarily equal part ) and then add up the results to get a 4 square representation. </p>\n\n<p><a href=\"https://math.stackexchange.com/questions/1972771/is-this-the-general-solution-of-finding-the-two-original-squares-that-add-up-to/1977767?noredirect=1#comment4062390_1977767\">https://math.stackexchange.com/questions/1972771/is-this-the-general-solution-of-finding-the-two-original-squares-that-add-up-to/1977767?noredirect=1#comment4062390_1977767</a></p>\n" } ]
Question: Efficient method to write number as a sum of four squares? Answer 1: <p>One of their methods for $n=4k+2$ is as follows:</p> <p>Randomly select an even number $a$ and an odd number $b$ such that $a^2+b^2 &lt; n$. Then, we hope $p=n-a^2-b^2$ is a prime (you can show there's about a $1/(A \log n \log \log n)$ chance of $p$ being prime ); $p$ is of the form $4r+1$, so if $p$ is prime there's a solution to $c^2+d^2=p$.</p> <p>To find that, we try to solve $m^2+1 \equiv 0 \pmod{p}$; I'm actually going to describe a slightly different method. Select $x$ at random from $1$ to $p-1$; then, $x^{2r}=\pm 1$ depending on whether $x$ is a quadratic residue so calculate $x^r$ by repeated squaring, and with a $1/2$ chance if $p$ is prime (a smaller one if $p$ is composite) you'll find a valid $m$.</p> <p>Given such an $m$, $m+i$ is a Gaussian integer with norm divisible by $p$ but smaller than $p^2$; use the Euclidean algorithm on the Gaussian integers with $p$ and get $c+di$ with norm $p$, and $a^2+b^2+c^2+d^2=n$.</p> <p>For an odd number $n$, solve for $a^2+b^2+c^2+d^2=2n$; note that by mod $4$ considerations exactly two of $a,b,c,d$ must be odd, and without loss of generality assume $a,b$ are odd and $c,d$ are even. Then,$(\frac{1}{2}(a+b))^2+(\frac{1}{2}(a-b))^2+(\frac{1}{2}(c+d))^2+(\frac{1}{2}(c-d))^2=n.$</p> <p>For $n$ a multiple of $4$, solve for $n/4$ recursively, and multiply all values by $4$.</p> Answer 2: <p>This seems to be addressed in <a href="https://pdfs.semanticscholar.org/9bd7/f2b063bfd2ec831be9c58750287d8a77b186.pdf" rel="nofollow noreferrer">this paper by Bumby.</a></p> Answer 3: <p>I am going to provide you with 4 links and I hope it will help you.</p> <p>this one is an online calculator based on an algorithm from math overflow.</p> <p><a href="http://www.mathcelebrity.com/foursquare.php?num=+178&amp;pl=Show+Lagrange+Four-Square+Notation" rel="nofollow noreferrer">http://www.mathcelebrity.com/foursquare.php?num=+178&amp;pl=Show+Lagrange+Four-Square+Notation</a> </p> <p>the math overflow algorithm can be found here.</p> <p><a href="https://stackoverflow.com/questions/11732555/how-to-find-all-possible-values-of-four-variables-when-squared-sum-to-n">https://stackoverflow.com/questions/11732555/how-to-find-all-possible-values-of-four-variables-when-squared-sum-to-n</a> </p> <p>and then there is also this algorithm from cs stackexchange </p> <p><a href="https://cs.stackexchange.com/questions/2988/how-fast-can-we-find-all-four-square-combinations-that-sum-to-n">https://cs.stackexchange.com/questions/2988/how-fast-can-we-find-all-four-square-combinations-that-sum-to-n</a></p> <p>and if you want to do a bit of work, then there is this algorithm that can find the sum of 2 squares for a given integer. So you would have to divide your number in two ( not necessarily equal part ) and then add up the results to get a 4 square representation. </p> <p><a href="https://math.stackexchange.com/questions/1972771/is-this-the-general-solution-of-finding-the-two-original-squares-that-add-up-to/1977767?noredirect=1#comment4062390_1977767">https://math.stackexchange.com/questions/1972771/is-this-the-general-solution-of-finding-the-two-original-squares-that-add-up-to/1977767?noredirect=1#comment4062390_1977767</a></p> Answer 4: <p>One of their methods for $n=4k+2$ is as follows:</p> <p>Randomly select an even number $a$ and an odd number $b$ such that $a^2+b^2 &lt; n$. Then, we hope $p=n-a^2-b^2$ is a prime (you can show there's about a $1/(A \log n \log \log n)$ chance of $p$ being prime ); $p$ is of the form $4r+1$, so if $p$ is prime there's a solution to $c^2+d^2=p$.</p> <p>To find that, we try to solve $m^2+1 \equiv 0 \pmod{p}$; I'm actually going to describe a slightly different method. Select $x$ at random from $1$ to $p-1$; then, $x^{2r}=\pm 1$ depending on whether $x$ is a quadratic residue so calculate $x^r$ by repeated squaring, and with a $1/2$ chance if $p$ is prime (a smaller one if $p$ is composite) you'll find a valid $m$.</p> <p>Given such an $m$, $m+i$ is a Gaussian integer with norm divisible by $p$ but smaller than $p^2$; use the Euclidean algorithm on the Gaussian integers with $p$ and get $c+di$ with norm $p$, and $a^2+b^2+c^2+d^2=n$.</p> <p>For an odd number $n$, solve for $a^2+b^2+c^2+d^2=2n$; note that by mod $4$ considerations exactly two of $a,b,c,d$ must be odd, and without loss of generality assume $a,b$ are odd and $c,d$ are even. Then,$(\frac{1}{2}(a+b))^2+(\frac{1}{2}(a-b))^2+(\frac{1}{2}(c+d))^2+(\frac{1}{2}(c-d))^2=n.$</p> <p>For $n$ a multiple of $4$, solve for $n/4$ recursively, and multiply all values by $4$.</p> Answer 5: <p>This seems to be addressed in <a href="https://pdfs.semanticscholar.org/9bd7/f2b063bfd2ec831be9c58750287d8a77b186.pdf" rel="nofollow noreferrer">this paper by Bumby.</a></p> Answer 6: <p>I am going to provide you with 4 links and I hope it will help you.</p> <p>this one is an online calculator based on an algorithm from math overflow.</p> <p><a href="http://www.mathcelebrity.com/foursquare.php?num=+178&amp;pl=Show+Lagrange+Four-Square+Notation" rel="nofollow noreferrer">http://www.mathcelebrity.com/foursquare.php?num=+178&amp;pl=Show+Lagrange+Four-Square+Notation</a> </p> <p>the math overflow algorithm can be found here.</p> <p><a href="https://stackoverflow.com/questions/11732555/how-to-find-all-possible-values-of-four-variables-when-squared-sum-to-n">https://stackoverflow.com/questions/11732555/how-to-find-all-possible-values-of-four-variables-when-squared-sum-to-n</a> </p> <p>and then there is also this algorithm from cs stackexchange </p> <p><a href="https://cs.stackexchange.com/questions/2988/how-fast-can-we-find-all-four-square-combinations-that-sum-to-n">https://cs.stackexchange.com/questions/2988/how-fast-can-we-find-all-four-square-combinations-that-sum-to-n</a></p> <p>and if you want to do a bit of work, then there is this algorithm that can find the sum of 2 squares for a given integer. So you would have to divide your number in two ( not necessarily equal part ) and then add up the results to get a 4 square representation. </p> <p><a href="https://math.stackexchange.com/questions/1972771/is-this-the-general-solution-of-finding-the-two-original-squares-that-add-up-to/1977767?noredirect=1#comment4062390_1977767">https://math.stackexchange.com/questions/1972771/is-this-the-general-solution-of-finding-the-two-original-squares-that-add-up-to/1977767?noredirect=1#comment4062390_1977767</a></p>
What is the density of integers of the form $a^2+nb^2$?
259,217
9
https://mathoverflow.net/questions/259217/what-is-the-density-of-integers-of-the-form-a2nb2
<p>Landau proved that the mean density of integers of the form $a^2+b^2$ up to $x$ is $K\frac{x}{\sqrt{\log x}} (1+o(1))$, where $K$ is an explicit constant. One proof is based on the fact that a prime $p$ is of this form iff $p\neq 3\mod 4$, and so we can express the generating function for these integers in terms of the zeta function and the $L$-function of the non-trivial character modulo $4$. From there one proceed using complex analysis. </p> <p>For integers of the form $a^2 + nb^2$, one should have similar results. We know what primes has this form by class field theory, and then in principle one can continue as before. </p> <p>I am sure this is classical, but unfortunately I've failed to find it in the literature (including google). <strong>Does anyone know a reference for this?</strong> </p>
[ { "answer_id": 259220, "score": 11, "text": "<p>Yes there is quite a bit in the literature on this problem. Apparently it was first solved by Bernays in his 1912 PhD thesis under Landau. The density is as expected (namely proportional to $x/\\sqrt{\\log x}.$)</p>\n\n<p>A modern paper on the topic is:</p>\n\n<p>Brink, Moree, Osburn - Principal forms $X^2+nY^2$ representing many integers.</p>\n\n<p>This contains many useful references about the history of the problem.</p>\n" }, { "answer_id": 259333, "score": 3, "text": "<p>R. W. K. Odoni has his own proof of the theorem, <a href=\"http://journals.cambridge.org/article_S0025579300005945\" rel=\"nofollow noreferrer\">see this paper on the problem</a>,</p>\n\n<blockquote>\n <p>\"On norms of integers in a full module of an algebraic number field and the distribution of values of binary integral quadratic forms\",</p>\n</blockquote>\n\n<p>published in Mathematika 22 (1975), no. 2, 108–111. See <a href=\"http://www.ams.org/mathscinet-getitem?mr=389856\" rel=\"nofollow noreferrer\">Narkiewicz's review</a>. Odoni's paper does not appear in the paper mentioned by Daniel.</p>\n\n<p>The concluding remarks in Odoni's paper are very interesting and refer to some earlier partial results.</p>\n" }, { "answer_id": 259220, "score": 11, "text": "<p>Yes there is quite a bit in the literature on this problem. Apparently it was first solved by Bernays in his 1912 PhD thesis under Landau. The density is as expected (namely proportional to $x/\\sqrt{\\log x}.$)</p>\n\n<p>A modern paper on the topic is:</p>\n\n<p>Brink, Moree, Osburn - Principal forms $X^2+nY^2$ representing many integers.</p>\n\n<p>This contains many useful references about the history of the problem.</p>\n" }, { "answer_id": 259333, "score": 3, "text": "<p>R. W. K. Odoni has his own proof of the theorem, <a href=\"http://journals.cambridge.org/article_S0025579300005945\" rel=\"nofollow noreferrer\">see this paper on the problem</a>,</p>\n\n<blockquote>\n <p>\"On norms of integers in a full module of an algebraic number field and the distribution of values of binary integral quadratic forms\",</p>\n</blockquote>\n\n<p>published in Mathematika 22 (1975), no. 2, 108–111. See <a href=\"http://www.ams.org/mathscinet-getitem?mr=389856\" rel=\"nofollow noreferrer\">Narkiewicz's review</a>. Odoni's paper does not appear in the paper mentioned by Daniel.</p>\n\n<p>The concluding remarks in Odoni's paper are very interesting and refer to some earlier partial results.</p>\n" } ]
Question: What is the density of integers of the form $a^2+nb^2$? Answer 1: <p>Yes there is quite a bit in the literature on this problem. Apparently it was first solved by Bernays in his 1912 PhD thesis under Landau. The density is as expected (namely proportional to $x/\sqrt{\log x}.$)</p> <p>A modern paper on the topic is:</p> <p>Brink, Moree, Osburn - Principal forms $X^2+nY^2$ representing many integers.</p> <p>This contains many useful references about the history of the problem.</p> Answer 2: <p>R. W. K. Odoni has his own proof of the theorem, <a href="http://journals.cambridge.org/article_S0025579300005945" rel="nofollow noreferrer">see this paper on the problem</a>,</p> <blockquote> <p>"On norms of integers in a full module of an algebraic number field and the distribution of values of binary integral quadratic forms",</p> </blockquote> <p>published in Mathematika 22 (1975), no. 2, 108–111. See <a href="http://www.ams.org/mathscinet-getitem?mr=389856" rel="nofollow noreferrer">Narkiewicz's review</a>. Odoni's paper does not appear in the paper mentioned by Daniel.</p> <p>The concluding remarks in Odoni's paper are very interesting and refer to some earlier partial results.</p> Answer 3: <p>Yes there is quite a bit in the literature on this problem. Apparently it was first solved by Bernays in his 1912 PhD thesis under Landau. The density is as expected (namely proportional to $x/\sqrt{\log x}.$)</p> <p>A modern paper on the topic is:</p> <p>Brink, Moree, Osburn - Principal forms $X^2+nY^2$ representing many integers.</p> <p>This contains many useful references about the history of the problem.</p> Answer 4: <p>R. W. K. Odoni has his own proof of the theorem, <a href="http://journals.cambridge.org/article_S0025579300005945" rel="nofollow noreferrer">see this paper on the problem</a>,</p> <blockquote> <p>"On norms of integers in a full module of an algebraic number field and the distribution of values of binary integral quadratic forms",</p> </blockquote> <p>published in Mathematika 22 (1975), no. 2, 108–111. See <a href="http://www.ams.org/mathscinet-getitem?mr=389856" rel="nofollow noreferrer">Narkiewicz's review</a>. Odoni's paper does not appear in the paper mentioned by Daniel.</p> <p>The concluding remarks in Odoni's paper are very interesting and refer to some earlier partial results.</p>
group actions in dimension 2 and 3
260,171
9
https://mathoverflow.net/questions/260171/group-actions-in-dimension-2-and-3
<p>I am looking for a reference to the following claims:</p> <ol> <li>Any compact group (connected or not) acting on $S^2$ is differentiably conjugate to a linear action. This must be classical. </li> <li>A circle $S^1$ acting on $RP^3$ (and supposedly any spherical space form) is differentiably conjugate to a linear action. This is probably true for every compact group acting on a $3$-dimensional spherical space form?</li> </ol> <p>Wolfgang Ziller</p>
[ { "answer_id": 260173, "score": 7, "text": "<p>For (2), see the paper of Frank Raymond (1968).</p>\n\n<p><em>Frank Raymond</em>, MR 219086 <a href=\"http://www.ams.org/mathscinet-getitem?mr=219086\"><strong>Classification of the actions of the circle on $3$-manifolds</strong></a>, <em>Trans. Amer. Math. Soc.</em> <strong>131</strong> (1968), 51--78.</p>\n\n<p>For (1) there is <a href=\"http://elib.uni-stuttgart.de/bitstream/11682/4652/1/ActStu.pdf\">the thesis of Harald Biller at Frankfurt</a> (1999).</p>\n" }, { "answer_id": 260200, "score": 7, "text": "<p>For (1), the usual trick is to average a Riemannian metric (using\nHaar measure on the group) to get a fixed metric. Hence the group\nwill be a subgroup of $SO(3)$ (e.g., by uniformization, the metric\non $S^2$ is conformal to the round $S^2$, so the group must act\nby conformal transformations, i.e. $PSL_2(\\mathbb{C})$. The\nmaximal compact subgroup of $PSL_2(\\mathbb{C})$ is $SO(3)$). </p>\n\n<p>For (2), the quotient space is a 2-dimensional orbifold (maybe\nwith boundary corresponding to the fixed locus). Away from the\nboundary, it will be a Seifert fibration, and one may use\nthe classification of Seifert fibrations to see that the action\nis conjugate to a linear action. </p>\n" }, { "answer_id": 260173, "score": 7, "text": "<p>For (2), see the paper of Frank Raymond (1968).</p>\n\n<p><em>Frank Raymond</em>, MR 219086 <a href=\"http://www.ams.org/mathscinet-getitem?mr=219086\"><strong>Classification of the actions of the circle on $3$-manifolds</strong></a>, <em>Trans. Amer. Math. Soc.</em> <strong>131</strong> (1968), 51--78.</p>\n\n<p>For (1) there is <a href=\"http://elib.uni-stuttgart.de/bitstream/11682/4652/1/ActStu.pdf\">the thesis of Harald Biller at Frankfurt</a> (1999).</p>\n" }, { "answer_id": 260200, "score": 7, "text": "<p>For (1), the usual trick is to average a Riemannian metric (using\nHaar measure on the group) to get a fixed metric. Hence the group\nwill be a subgroup of $SO(3)$ (e.g., by uniformization, the metric\non $S^2$ is conformal to the round $S^2$, so the group must act\nby conformal transformations, i.e. $PSL_2(\\mathbb{C})$. The\nmaximal compact subgroup of $PSL_2(\\mathbb{C})$ is $SO(3)$). </p>\n\n<p>For (2), the quotient space is a 2-dimensional orbifold (maybe\nwith boundary corresponding to the fixed locus). Away from the\nboundary, it will be a Seifert fibration, and one may use\nthe classification of Seifert fibrations to see that the action\nis conjugate to a linear action. </p>\n" } ]
Question: group actions in dimension 2 and 3 Answer 1: <p>For (2), see the paper of Frank Raymond (1968).</p> <p><em>Frank Raymond</em>, MR 219086 <a href="http://www.ams.org/mathscinet-getitem?mr=219086"><strong>Classification of the actions of the circle on $3$-manifolds</strong></a>, <em>Trans. Amer. Math. Soc.</em> <strong>131</strong> (1968), 51--78.</p> <p>For (1) there is <a href="http://elib.uni-stuttgart.de/bitstream/11682/4652/1/ActStu.pdf">the thesis of Harald Biller at Frankfurt</a> (1999).</p> Answer 2: <p>For (1), the usual trick is to average a Riemannian metric (using Haar measure on the group) to get a fixed metric. Hence the group will be a subgroup of $SO(3)$ (e.g., by uniformization, the metric on $S^2$ is conformal to the round $S^2$, so the group must act by conformal transformations, i.e. $PSL_2(\mathbb{C})$. The maximal compact subgroup of $PSL_2(\mathbb{C})$ is $SO(3)$). </p> <p>For (2), the quotient space is a 2-dimensional orbifold (maybe with boundary corresponding to the fixed locus). Away from the boundary, it will be a Seifert fibration, and one may use the classification of Seifert fibrations to see that the action is conjugate to a linear action. </p> Answer 3: <p>For (2), see the paper of Frank Raymond (1968).</p> <p><em>Frank Raymond</em>, MR 219086 <a href="http://www.ams.org/mathscinet-getitem?mr=219086"><strong>Classification of the actions of the circle on $3$-manifolds</strong></a>, <em>Trans. Amer. Math. Soc.</em> <strong>131</strong> (1968), 51--78.</p> <p>For (1) there is <a href="http://elib.uni-stuttgart.de/bitstream/11682/4652/1/ActStu.pdf">the thesis of Harald Biller at Frankfurt</a> (1999).</p> Answer 4: <p>For (1), the usual trick is to average a Riemannian metric (using Haar measure on the group) to get a fixed metric. Hence the group will be a subgroup of $SO(3)$ (e.g., by uniformization, the metric on $S^2$ is conformal to the round $S^2$, so the group must act by conformal transformations, i.e. $PSL_2(\mathbb{C})$. The maximal compact subgroup of $PSL_2(\mathbb{C})$ is $SO(3)$). </p> <p>For (2), the quotient space is a 2-dimensional orbifold (maybe with boundary corresponding to the fixed locus). Away from the boundary, it will be a Seifert fibration, and one may use the classification of Seifert fibrations to see that the action is conjugate to a linear action. </p>
If normal with respect to prime base then normal for all bases
260,442
9
https://mathoverflow.net/questions/260442/if-normal-with-respect-to-prime-base-then-normal-for-all-bases
<p>I tried to find it on internet but couldn't so m asking this here. I want to ask if a number is normal with respect to all prime number base then do we know that it is normal with respect to any base. Obviously we know that if a number is normal with respect to base b then for any $ b^k$ it is normal. </p> <p>Thanks</p>
[ { "answer_id": 260674, "score": 9, "text": "<p>Gerry Myerson already answered in the comments, but let me be a little more explicit. Take any number $m\\in\\mathbb{N}$ with at least two prime factors (for example $6$). Then construct a Cantor set $C=C_m$ similar to the ternary Cantor set, except that instead of keeping two intervals of relative length $1/3$ we keep two intervals of relative length $1/m$ at each step. Alternatively, $C_m$ is the set of points in $[0,1]$ whose base $m$ expansion has only digits $0, m-1$. Clearly, no point in $C$ is remotely normal to base $m$.</p>\n\n<p>On the other hand, $C$ carries a natural measure $\\mu$, which can be defined by the property that is assigns the same mass $2^{-k}$ to all $2^k$ intervals in the $k$-th stage of the construction ($\\mu$ is also a multiple of Hausdorff measure of the appropriate dimension on $C$). </p>\n\n<p>It is then known that $\\mu$-almost all points are normal to any base $p$ such that $\\log p/\\log m$ is irrational. In particular, this is true for all prime numbers $p$. So $\\mu$-almost all points are counterexamples to the question.</p>\n\n<p>The proof of this essentially goes back to Cassels [<strong>Cassels, J. W. S.</strong> <a href=\"https://eudml.org/doc/210691\" rel=\"nofollow noreferrer\">On a problem of Steinhaus about normal numbers.</a> <em>Colloq. Math.</em> 7 1959 95--101], who proved the special case $m=3$, by looking at the Fourier transform of $\\mu$ along sequences $p^n \\xi$. I am quite sure the proof works for any $m$. See also our paper [<strong>Hochman, Michael; Shmerkin, Pablo.</strong> <a href=\"https://doi.org/10.1007/s00222-014-0573-5\" rel=\"nofollow noreferrer\">Equidistribution from fractal measures.</a> <em>Invent. Math.</em> 202 (2015), no. 1, 427--479.] for more discussion and generalizations.</p>\n" }, { "answer_id": 260676, "score": 7, "text": "<p>Wolfgang Schmidt, <a href=\"https://eudml.org/doc/207013\" rel=\"nofollow noreferrer\">Über die Normalität von Zahlen zu verschiedenen Basen</a>, Acta Arith. 7 1961/1962 299–309, <a href=\"http://www.ams.org/mathscinet-getitem?mr=0140482\" rel=\"nofollow noreferrer\">MR0140482</a> (25 #3902), according to the review by N. G. de Bruijn, proved this: Let the set of all integers $&gt;1$ be split into two disjoint classes $R$ and $S$, such that for every $n&gt;1$ all powers of $n$ belong to the same class as $n$ itself. Then there exists a continuum of numbers $\\xi$ which all have the property that $\\xi$ is normal with respect to every $r\\in R$ and non-normal with respect to every $s\\in S$. </p>\n\n<p>In particular, let $R$ be the set of all primes and prime powers, $S$ the set of all integers with two or more distinct prime factors. Then there is a continuum (that is, an uncountable infinity) of numbers normal to every prime base and not normal to any base that isn't prime. </p>\n" }, { "answer_id": 260674, "score": 9, "text": "<p>Gerry Myerson already answered in the comments, but let me be a little more explicit. Take any number $m\\in\\mathbb{N}$ with at least two prime factors (for example $6$). Then construct a Cantor set $C=C_m$ similar to the ternary Cantor set, except that instead of keeping two intervals of relative length $1/3$ we keep two intervals of relative length $1/m$ at each step. Alternatively, $C_m$ is the set of points in $[0,1]$ whose base $m$ expansion has only digits $0, m-1$. Clearly, no point in $C$ is remotely normal to base $m$.</p>\n\n<p>On the other hand, $C$ carries a natural measure $\\mu$, which can be defined by the property that is assigns the same mass $2^{-k}$ to all $2^k$ intervals in the $k$-th stage of the construction ($\\mu$ is also a multiple of Hausdorff measure of the appropriate dimension on $C$). </p>\n\n<p>It is then known that $\\mu$-almost all points are normal to any base $p$ such that $\\log p/\\log m$ is irrational. In particular, this is true for all prime numbers $p$. So $\\mu$-almost all points are counterexamples to the question.</p>\n\n<p>The proof of this essentially goes back to Cassels [<strong>Cassels, J. W. S.</strong> <a href=\"https://eudml.org/doc/210691\" rel=\"nofollow noreferrer\">On a problem of Steinhaus about normal numbers.</a> <em>Colloq. Math.</em> 7 1959 95--101], who proved the special case $m=3$, by looking at the Fourier transform of $\\mu$ along sequences $p^n \\xi$. I am quite sure the proof works for any $m$. See also our paper [<strong>Hochman, Michael; Shmerkin, Pablo.</strong> <a href=\"https://doi.org/10.1007/s00222-014-0573-5\" rel=\"nofollow noreferrer\">Equidistribution from fractal measures.</a> <em>Invent. Math.</em> 202 (2015), no. 1, 427--479.] for more discussion and generalizations.</p>\n" }, { "answer_id": 260676, "score": 7, "text": "<p>Wolfgang Schmidt, <a href=\"https://eudml.org/doc/207013\" rel=\"nofollow noreferrer\">Über die Normalität von Zahlen zu verschiedenen Basen</a>, Acta Arith. 7 1961/1962 299–309, <a href=\"http://www.ams.org/mathscinet-getitem?mr=0140482\" rel=\"nofollow noreferrer\">MR0140482</a> (25 #3902), according to the review by N. G. de Bruijn, proved this: Let the set of all integers $&gt;1$ be split into two disjoint classes $R$ and $S$, such that for every $n&gt;1$ all powers of $n$ belong to the same class as $n$ itself. Then there exists a continuum of numbers $\\xi$ which all have the property that $\\xi$ is normal with respect to every $r\\in R$ and non-normal with respect to every $s\\in S$. </p>\n\n<p>In particular, let $R$ be the set of all primes and prime powers, $S$ the set of all integers with two or more distinct prime factors. Then there is a continuum (that is, an uncountable infinity) of numbers normal to every prime base and not normal to any base that isn't prime. </p>\n" } ]
Question: If normal with respect to prime base then normal for all bases Answer 1: <p>Gerry Myerson already answered in the comments, but let me be a little more explicit. Take any number $m\in\mathbb{N}$ with at least two prime factors (for example $6$). Then construct a Cantor set $C=C_m$ similar to the ternary Cantor set, except that instead of keeping two intervals of relative length $1/3$ we keep two intervals of relative length $1/m$ at each step. Alternatively, $C_m$ is the set of points in $[0,1]$ whose base $m$ expansion has only digits $0, m-1$. Clearly, no point in $C$ is remotely normal to base $m$.</p> <p>On the other hand, $C$ carries a natural measure $\mu$, which can be defined by the property that is assigns the same mass $2^{-k}$ to all $2^k$ intervals in the $k$-th stage of the construction ($\mu$ is also a multiple of Hausdorff measure of the appropriate dimension on $C$). </p> <p>It is then known that $\mu$-almost all points are normal to any base $p$ such that $\log p/\log m$ is irrational. In particular, this is true for all prime numbers $p$. So $\mu$-almost all points are counterexamples to the question.</p> <p>The proof of this essentially goes back to Cassels [<strong>Cassels, J. W. S.</strong> <a href="https://eudml.org/doc/210691" rel="nofollow noreferrer">On a problem of Steinhaus about normal numbers.</a> <em>Colloq. Math.</em> 7 1959 95--101], who proved the special case $m=3$, by looking at the Fourier transform of $\mu$ along sequences $p^n \xi$. I am quite sure the proof works for any $m$. See also our paper [<strong>Hochman, Michael; Shmerkin, Pablo.</strong> <a href="https://doi.org/10.1007/s00222-014-0573-5" rel="nofollow noreferrer">Equidistribution from fractal measures.</a> <em>Invent. Math.</em> 202 (2015), no. 1, 427--479.] for more discussion and generalizations.</p> Answer 2: <p>Wolfgang Schmidt, <a href="https://eudml.org/doc/207013" rel="nofollow noreferrer">Über die Normalität von Zahlen zu verschiedenen Basen</a>, Acta Arith. 7 1961/1962 299–309, <a href="http://www.ams.org/mathscinet-getitem?mr=0140482" rel="nofollow noreferrer">MR0140482</a> (25 #3902), according to the review by N. G. de Bruijn, proved this: Let the set of all integers $&gt;1$ be split into two disjoint classes $R$ and $S$, such that for every $n&gt;1$ all powers of $n$ belong to the same class as $n$ itself. Then there exists a continuum of numbers $\xi$ which all have the property that $\xi$ is normal with respect to every $r\in R$ and non-normal with respect to every $s\in S$. </p> <p>In particular, let $R$ be the set of all primes and prime powers, $S$ the set of all integers with two or more distinct prime factors. Then there is a continuum (that is, an uncountable infinity) of numbers normal to every prime base and not normal to any base that isn't prime. </p> Answer 3: <p>Gerry Myerson already answered in the comments, but let me be a little more explicit. Take any number $m\in\mathbb{N}$ with at least two prime factors (for example $6$). Then construct a Cantor set $C=C_m$ similar to the ternary Cantor set, except that instead of keeping two intervals of relative length $1/3$ we keep two intervals of relative length $1/m$ at each step. Alternatively, $C_m$ is the set of points in $[0,1]$ whose base $m$ expansion has only digits $0, m-1$. Clearly, no point in $C$ is remotely normal to base $m$.</p> <p>On the other hand, $C$ carries a natural measure $\mu$, which can be defined by the property that is assigns the same mass $2^{-k}$ to all $2^k$ intervals in the $k$-th stage of the construction ($\mu$ is also a multiple of Hausdorff measure of the appropriate dimension on $C$). </p> <p>It is then known that $\mu$-almost all points are normal to any base $p$ such that $\log p/\log m$ is irrational. In particular, this is true for all prime numbers $p$. So $\mu$-almost all points are counterexamples to the question.</p> <p>The proof of this essentially goes back to Cassels [<strong>Cassels, J. W. S.</strong> <a href="https://eudml.org/doc/210691" rel="nofollow noreferrer">On a problem of Steinhaus about normal numbers.</a> <em>Colloq. Math.</em> 7 1959 95--101], who proved the special case $m=3$, by looking at the Fourier transform of $\mu$ along sequences $p^n \xi$. I am quite sure the proof works for any $m$. See also our paper [<strong>Hochman, Michael; Shmerkin, Pablo.</strong> <a href="https://doi.org/10.1007/s00222-014-0573-5" rel="nofollow noreferrer">Equidistribution from fractal measures.</a> <em>Invent. Math.</em> 202 (2015), no. 1, 427--479.] for more discussion and generalizations.</p> Answer 4: <p>Wolfgang Schmidt, <a href="https://eudml.org/doc/207013" rel="nofollow noreferrer">Über die Normalität von Zahlen zu verschiedenen Basen</a>, Acta Arith. 7 1961/1962 299–309, <a href="http://www.ams.org/mathscinet-getitem?mr=0140482" rel="nofollow noreferrer">MR0140482</a> (25 #3902), according to the review by N. G. de Bruijn, proved this: Let the set of all integers $&gt;1$ be split into two disjoint classes $R$ and $S$, such that for every $n&gt;1$ all powers of $n$ belong to the same class as $n$ itself. Then there exists a continuum of numbers $\xi$ which all have the property that $\xi$ is normal with respect to every $r\in R$ and non-normal with respect to every $s\in S$. </p> <p>In particular, let $R$ be the set of all primes and prime powers, $S$ the set of all integers with two or more distinct prime factors. Then there is a continuum (that is, an uncountable infinity) of numbers normal to every prime base and not normal to any base that isn't prime. </p>
Classification of $p$-groups, what after it?
260,673
9
https://mathoverflow.net/questions/260673/classification-of-p-groups-what-after-it
<p>In finite group theory, $p$-groups or simple groups can be considered as building blocks of all the groups. What is known about these families of groups is that </p> <ol> <li><p>The classification of simple groups is done.</p></li> <li><p>The classification of $p$-groups <em>is much difficult</em> problem.</p></li> </ol> <p>Due to classification of simple groups, many problems from different areas of mathematics have been solved; there is a book <em>Atlas of Finite Simple Groups: Ten years on</em>, which perhaps contains some work with this thought.</p> <p>On the other hand, although classification of $p$-groups is difficult, many researchers are focused on <em>classifying $p$-groups having a specific property</em>. For example, such work is collected in five big volumes (each $\geq$ 600 page, multiple of thousands of problems in total), only devoted to finite $p$-groups by Berkovich and Janko, in which, the authors state in preface that their aim is classification w.r.t. some properties.</p> <p>I wondered many times, but didn't get any definite answer to following question while discussing with some people working on $p$-groups. I thought it may not be good to post this question on group-pub-forum to de-motivate some group-theoriests. The question is </p> <blockquote> <p>As long as <em>classification of some types of $p$-groups</em> is concerned, has its application appeared in some other branch of mathematics (or at least, in algebra, or at least in group theory itself!)?</p> </blockquote>
[ { "answer_id": 261035, "score": 8, "text": "<p>Understanding the structure of various classes of $p$-groups already plays a substantial role in the classification of finite simple groups. To give (only a few of many possible), understanding $p$-groups in which all characteristic Abelian subgroups are cyclic ( so called $p$-groups of symplectic type), which were classified by P. Hall, is important in local analysis used for classification of simple groups. Understanding $p$-groups in which the graph with vertices elementary Abelian subgroups of order $p^{2}$ is disconnected, where two such subgroups are joined if they commute (elementwise) is relevant for the proof of the odd order theorem, and in signalizer functor theory elsewhere. Understanding $2$-groups of sectional rank at most $4$ is useful for CFSG. There are many other examples, and maybe more will occur during the ongoing revision projects for CFSG.</p>\n" }, { "answer_id": 296063, "score": 3, "text": "<p>Let me speak to part of your question: What after it? We can not reasonably expect to classify the $p^{2n^3/27 +O(n^{8/3})}$ groups of order $p^n$ for large $n$. We can classify $p$-groups with particular properties focusing on those with applications to other problems (as Geoff and Alireza mention). Alternatively, we can focus on\nunderstanding small and highly symmetric examples. Certainly in graph theory small and highly symmetric examples have played an important role e.g. $s$-arc transitive graphs (with $2\\leqslant s\\leqslant 7$) have applications to incidence geometry and abstract polytopes. This philosophy was behind our paper\n<a href=\"https://arxiv.org/abs/1603.05384\" rel=\"nofollow noreferrer\">Maximal linear groups induced on the Frattini quotient of a p-group</a>.\nIndeed, one could argue that most large $p$-groups with small automorphism groups are uninteresting in the same way that most large graphs with few symmetries are uninteresting.</p>\n" }, { "answer_id": 260757, "score": 2, "text": "<p>The answer is positive: since one must give a classification of at least one types of finite $p$-groups which I suggest to consider the classification of finite $p$-groups having a maximal subgroup which is cyclic; and next one must show us an application in group theory as it is mentioned above.</p>\n\n<p>For the latter one may look at the following book and then count the number of usage of the above classification (Theorem 1.2, page 22) to see how many application the above classification may have at least. Note that \"Application of an Application of the classification\" must be considered as \"an Application of the classification\". </p>\n\n<p>Berkovich, Yakov Groups of prime power order. Vol. 1. With a foreword by Zvonimir Janko. De Gruyter Expositions in Mathematics, 46. Walter de Gruyter GmbH &amp; Co. KG, Berlin, 2008. xx+512 pp. ISBN: 978-3-11-020418-6</p>\n\n<p>For example, the classification of non-abelian finite 2-groups $G$ with $|G:G'|=4$ (Proposition 1.6, page 26) follows from the classification of groups with a cyclic maximal subgroups.</p>\n\n<p>Another application of finite $p$-groups which is every time one use is the classification of groups of order at most $p^3$. This has certainly many applications in group theory, one can certainly find such applications in the following book: </p>\n\n<p>Leedham-Green, C. R.; McKay, S. The structure of groups of prime power order. London Mathematical Society Monographs. New Series, 27. Oxford Science Publications. Oxford University Press, Oxford, 2002. xii+334 pp. ISBN: 0-19-853548-1</p>\n\n<p>Let me finish my answer by saying that the ultimate aim of every branch of mathematics is to classify all objects which are studied therein; however the latter aim may not be possible, where the impossibility must be defined among the branch.</p>\n" }, { "answer_id": 261035, "score": 8, "text": "<p>Understanding the structure of various classes of $p$-groups already plays a substantial role in the classification of finite simple groups. To give (only a few of many possible), understanding $p$-groups in which all characteristic Abelian subgroups are cyclic ( so called $p$-groups of symplectic type), which were classified by P. Hall, is important in local analysis used for classification of simple groups. Understanding $p$-groups in which the graph with vertices elementary Abelian subgroups of order $p^{2}$ is disconnected, where two such subgroups are joined if they commute (elementwise) is relevant for the proof of the odd order theorem, and in signalizer functor theory elsewhere. Understanding $2$-groups of sectional rank at most $4$ is useful for CFSG. There are many other examples, and maybe more will occur during the ongoing revision projects for CFSG.</p>\n" }, { "answer_id": 296063, "score": 3, "text": "<p>Let me speak to part of your question: What after it? We can not reasonably expect to classify the $p^{2n^3/27 +O(n^{8/3})}$ groups of order $p^n$ for large $n$. We can classify $p$-groups with particular properties focusing on those with applications to other problems (as Geoff and Alireza mention). Alternatively, we can focus on\nunderstanding small and highly symmetric examples. Certainly in graph theory small and highly symmetric examples have played an important role e.g. $s$-arc transitive graphs (with $2\\leqslant s\\leqslant 7$) have applications to incidence geometry and abstract polytopes. This philosophy was behind our paper\n<a href=\"https://arxiv.org/abs/1603.05384\" rel=\"nofollow noreferrer\">Maximal linear groups induced on the Frattini quotient of a p-group</a>.\nIndeed, one could argue that most large $p$-groups with small automorphism groups are uninteresting in the same way that most large graphs with few symmetries are uninteresting.</p>\n" }, { "answer_id": 260757, "score": 2, "text": "<p>The answer is positive: since one must give a classification of at least one types of finite $p$-groups which I suggest to consider the classification of finite $p$-groups having a maximal subgroup which is cyclic; and next one must show us an application in group theory as it is mentioned above.</p>\n\n<p>For the latter one may look at the following book and then count the number of usage of the above classification (Theorem 1.2, page 22) to see how many application the above classification may have at least. Note that \"Application of an Application of the classification\" must be considered as \"an Application of the classification\". </p>\n\n<p>Berkovich, Yakov Groups of prime power order. Vol. 1. With a foreword by Zvonimir Janko. De Gruyter Expositions in Mathematics, 46. Walter de Gruyter GmbH &amp; Co. KG, Berlin, 2008. xx+512 pp. ISBN: 978-3-11-020418-6</p>\n\n<p>For example, the classification of non-abelian finite 2-groups $G$ with $|G:G'|=4$ (Proposition 1.6, page 26) follows from the classification of groups with a cyclic maximal subgroups.</p>\n\n<p>Another application of finite $p$-groups which is every time one use is the classification of groups of order at most $p^3$. This has certainly many applications in group theory, one can certainly find such applications in the following book: </p>\n\n<p>Leedham-Green, C. R.; McKay, S. The structure of groups of prime power order. London Mathematical Society Monographs. New Series, 27. Oxford Science Publications. Oxford University Press, Oxford, 2002. xii+334 pp. ISBN: 0-19-853548-1</p>\n\n<p>Let me finish my answer by saying that the ultimate aim of every branch of mathematics is to classify all objects which are studied therein; however the latter aim may not be possible, where the impossibility must be defined among the branch.</p>\n" } ]
Question: Classification of $p$-groups, what after it? Answer 1: <p>Understanding the structure of various classes of $p$-groups already plays a substantial role in the classification of finite simple groups. To give (only a few of many possible), understanding $p$-groups in which all characteristic Abelian subgroups are cyclic ( so called $p$-groups of symplectic type), which were classified by P. Hall, is important in local analysis used for classification of simple groups. Understanding $p$-groups in which the graph with vertices elementary Abelian subgroups of order $p^{2}$ is disconnected, where two such subgroups are joined if they commute (elementwise) is relevant for the proof of the odd order theorem, and in signalizer functor theory elsewhere. Understanding $2$-groups of sectional rank at most $4$ is useful for CFSG. There are many other examples, and maybe more will occur during the ongoing revision projects for CFSG.</p> Answer 2: <p>Let me speak to part of your question: What after it? We can not reasonably expect to classify the $p^{2n^3/27 +O(n^{8/3})}$ groups of order $p^n$ for large $n$. We can classify $p$-groups with particular properties focusing on those with applications to other problems (as Geoff and Alireza mention). Alternatively, we can focus on understanding small and highly symmetric examples. Certainly in graph theory small and highly symmetric examples have played an important role e.g. $s$-arc transitive graphs (with $2\leqslant s\leqslant 7$) have applications to incidence geometry and abstract polytopes. This philosophy was behind our paper <a href="https://arxiv.org/abs/1603.05384" rel="nofollow noreferrer">Maximal linear groups induced on the Frattini quotient of a p-group</a>. Indeed, one could argue that most large $p$-groups with small automorphism groups are uninteresting in the same way that most large graphs with few symmetries are uninteresting.</p> Answer 3: <p>The answer is positive: since one must give a classification of at least one types of finite $p$-groups which I suggest to consider the classification of finite $p$-groups having a maximal subgroup which is cyclic; and next one must show us an application in group theory as it is mentioned above.</p> <p>For the latter one may look at the following book and then count the number of usage of the above classification (Theorem 1.2, page 22) to see how many application the above classification may have at least. Note that "Application of an Application of the classification" must be considered as "an Application of the classification". </p> <p>Berkovich, Yakov Groups of prime power order. Vol. 1. With a foreword by Zvonimir Janko. De Gruyter Expositions in Mathematics, 46. Walter de Gruyter GmbH &amp; Co. KG, Berlin, 2008. xx+512 pp. ISBN: 978-3-11-020418-6</p> <p>For example, the classification of non-abelian finite 2-groups $G$ with $|G:G'|=4$ (Proposition 1.6, page 26) follows from the classification of groups with a cyclic maximal subgroups.</p> <p>Another application of finite $p$-groups which is every time one use is the classification of groups of order at most $p^3$. This has certainly many applications in group theory, one can certainly find such applications in the following book: </p> <p>Leedham-Green, C. R.; McKay, S. The structure of groups of prime power order. London Mathematical Society Monographs. New Series, 27. Oxford Science Publications. Oxford University Press, Oxford, 2002. xii+334 pp. ISBN: 0-19-853548-1</p> <p>Let me finish my answer by saying that the ultimate aim of every branch of mathematics is to classify all objects which are studied therein; however the latter aim may not be possible, where the impossibility must be defined among the branch.</p> Answer 4: <p>Understanding the structure of various classes of $p$-groups already plays a substantial role in the classification of finite simple groups. To give (only a few of many possible), understanding $p$-groups in which all characteristic Abelian subgroups are cyclic ( so called $p$-groups of symplectic type), which were classified by P. Hall, is important in local analysis used for classification of simple groups. Understanding $p$-groups in which the graph with vertices elementary Abelian subgroups of order $p^{2}$ is disconnected, where two such subgroups are joined if they commute (elementwise) is relevant for the proof of the odd order theorem, and in signalizer functor theory elsewhere. Understanding $2$-groups of sectional rank at most $4$ is useful for CFSG. There are many other examples, and maybe more will occur during the ongoing revision projects for CFSG.</p> Answer 5: <p>Let me speak to part of your question: What after it? We can not reasonably expect to classify the $p^{2n^3/27 +O(n^{8/3})}$ groups of order $p^n$ for large $n$. We can classify $p$-groups with particular properties focusing on those with applications to other problems (as Geoff and Alireza mention). Alternatively, we can focus on understanding small and highly symmetric examples. Certainly in graph theory small and highly symmetric examples have played an important role e.g. $s$-arc transitive graphs (with $2\leqslant s\leqslant 7$) have applications to incidence geometry and abstract polytopes. This philosophy was behind our paper <a href="https://arxiv.org/abs/1603.05384" rel="nofollow noreferrer">Maximal linear groups induced on the Frattini quotient of a p-group</a>. Indeed, one could argue that most large $p$-groups with small automorphism groups are uninteresting in the same way that most large graphs with few symmetries are uninteresting.</p> Answer 6: <p>The answer is positive: since one must give a classification of at least one types of finite $p$-groups which I suggest to consider the classification of finite $p$-groups having a maximal subgroup which is cyclic; and next one must show us an application in group theory as it is mentioned above.</p> <p>For the latter one may look at the following book and then count the number of usage of the above classification (Theorem 1.2, page 22) to see how many application the above classification may have at least. Note that "Application of an Application of the classification" must be considered as "an Application of the classification". </p> <p>Berkovich, Yakov Groups of prime power order. Vol. 1. With a foreword by Zvonimir Janko. De Gruyter Expositions in Mathematics, 46. Walter de Gruyter GmbH &amp; Co. KG, Berlin, 2008. xx+512 pp. ISBN: 978-3-11-020418-6</p> <p>For example, the classification of non-abelian finite 2-groups $G$ with $|G:G'|=4$ (Proposition 1.6, page 26) follows from the classification of groups with a cyclic maximal subgroups.</p> <p>Another application of finite $p$-groups which is every time one use is the classification of groups of order at most $p^3$. This has certainly many applications in group theory, one can certainly find such applications in the following book: </p> <p>Leedham-Green, C. R.; McKay, S. The structure of groups of prime power order. London Mathematical Society Monographs. New Series, 27. Oxford Science Publications. Oxford University Press, Oxford, 2002. xii+334 pp. ISBN: 0-19-853548-1</p> <p>Let me finish my answer by saying that the ultimate aim of every branch of mathematics is to classify all objects which are studied therein; however the latter aim may not be possible, where the impossibility must be defined among the branch.</p>
Quotient-free monoidal categories
258,535
8
https://mathoverflow.net/questions/258535/quotient-free-monoidal-categories
<p>Let $(\mathsf{Rel},\otimes,1)$ denote the monoidal category of sets and relations, where $1$ is the one-element set. I once conjectured (with a little help from Jamie Vicary) that $\mathsf{Rel}$ is "quotient-free" in the sense that if a strong monoidal functor $F\colon\mathsf{Rel}\to S$ identifies <em>any</em> parallel pair of morphisms, then $F$ identifies <em>every</em> parallel pair of morphisms, and hence it factors through the terminal monoidal category (since $\mathsf{Rel}$ has a zero-object). [I'd be happy to hear suggestions for a better name than "quotient-free monoidal category", or for a better way of thinking of such things.]</p> <p><strong>Definition:</strong> We say a monoidal category $M$ is <em>quotient-free</em> if for any monoidal category $S$ and strong monoidal functor $F\colon M\to S$, if $F(f_0)=F(f_1)$ for distinct morphisms $f_0\neq f_1\colon A\to B$ then $F$ factors through a terminal monoidal category.</p> <p>Explaining the conjecture to Tobias Fritz, he quickly proved it (by contradiction) as follows. </p> <p><strong>Proposition:</strong> The monoidal category $(\mathsf{Rel},\otimes,1)$ is quotient-free.</p> <p><strong>Proof (Fritz):</strong> Suppose that $A$ and $B$ are sets and that $R_0,R_1\colon A\to B$ are relations such that $R_0\neq R_1$. Then there exists $a\in A$ and $b\in B$ such that $(a,b)\notin R_0$ and $(a,b)\in R_1$ (without loss of generality). </p> <p>Let $e_a\colon 1\to A$ and $e_b\colon 1\to B$ correspond to the relations characterizing the subsets $\{a\}\subseteq A$ and $\{b\}\subseteq B$, respectively, and let $e'_b\colon B\to 1$ be the transpose of $e_b$. Then we have two <em>different</em> relations $$1\xrightarrow{e_a}A\xrightarrow{R_0\ ,\ R_1}B\xrightarrow{e'_b}1.$$ These ($e'_bR_0e_a$ and $e'_bR_1e_a$) are the only two relations $1\to 1$, equaling the "null" relation $\emptyset_{1,1}$ and the identity $\mathrm{id}_1$, respectively.</p> <p>Assuming now that $F(R_0)=F(R_1)$, we have $F(\mathrm{id}_1)=F(\emptyset_{1,1})$. It follows that $F$ identifies any given relation $X\colon C\to D$ with the null relation $\emptyset_{C,D}\colon C\to D$, because $$ FX\cong F(X)\otimes F(\mathrm{id}_1)= F(X)\otimes F(\emptyset_{1,1})\cong F(X\otimes\emptyset_{1,1})= F(\emptyset_{C,D}). $$ Thus for any set $A$, we obtain an isomorphism $F(A)\cong F(\emptyset)$, where $\emptyset$ is the zero-object of $\mathsf{Rel}$. $\square$</p> <p><strong>Question:</strong> What are other examples of quotient-free monoidal categories?</p> <p><strong>Question:</strong> Might we consider quotient-free monoidal categories as acting like fields, which are also somehow quotient-free? That is, maps to quotient-free monoidal categories would be analogous to points? Any thoughts on this would be useful.</p>
[ { "answer_id": 258644, "score": 4, "text": "<p>Here are some additional thoughts and examples.</p>\n\n<p><strong>1)</strong> The argument presented in the OP applies whenever one has a monoidal category $(\\mathsf{C},\\otimes,1)$ in which the unit object $1$ is a <a href=\"https://ncatlab.org/nlab/show/separator\" rel=\"nofollow noreferrer\">separator</a> and coseparator, and the monoid of scalars $\\mathsf{C}(1,1)$ is itself quotient-free. So under these assumptions, also $\\mathsf{C}$ is quotient-free as a monoidal category.</p>\n\n<p><strong>2)</strong> A similar argument shows that for any field $K\\neq\\mathbb{Z}/2$, the category $\\mathsf{Vect}_K$ has exactly one nontrivial monoidal quotient, namely a category that I will denote by $\\mathsf{Vect}_K/K^\\times$. This quotient is given by identifying any two parallel morphisms that are scalar multiples of each other. (For $K=\\mathbb{Z}/2$, this quotient coincides with the original category, and already $\\mathsf{Vect}_{\\mathbb{Z}/2}$ itself is quotient-free by the same token.)</p>\n\n<p>To prove this claim, note again that the monoidal unit $K\\in\\mathsf{Vect}_K$ is a separator and coseparator, so that any nontrivial quotient of $\\mathsf{Vect}_K$ must also identify two different scalars. But the scalars are the multiplicative monoid $(K,\\cdot)$, which consists of the group of units $(K^\\times,\\cdot)$ together with an additional absorbing element $0\\in K$. Its only nontrivial quotient is the one where all units are identified, with quotient $(\\mathbb{Z}/2,\\cdot)$, which is accidentally also the monoid of scalars of $\\mathsf{Rel}$. By strong monoidality, our original quotient of $\\mathsf{Vect}_K$ must therefore also be a quotient of $\\mathsf{Vect}_K/K^\\times$.</p>\n\n<p>Finally, it remains to be shown that $\\mathsf{Vect}_K/K^\\times$ is quotient-free. So suppose that we have a quotient that identifies two parallel morphisms $[f],[g]:V\\to W$, where I'm working with representatives $f,g\\in\\mathsf{Vect}_K(V,W)$. By suitably pre- and postcomposing with a subspace inclusion and quotient space projection, we may assume that $f=\\mathrm{id}_V$, while $g$ is not a scalar multiple of the identity. In particular, we can choose a basis in which $g$ is not diagonal, resulting in an element $x:K\\to V$ and a functional $t:V\\to K$ such that $tx = tfx =0$, but $tgx\\neq 0$. Hence the quotient must identify the two remaining scalars $0,1\\in\\mathsf{Vect}_K/K^\\times(K,K)$, and we are again done.</p>\n\n<p><strong>3)</strong> The same statement and proof apply to the category $\\mathsf{FVect}_K$ of finite-dimensional vector spaces. But this category is also the free additive $K$-linear category, which makes it quotient-free in a different sense, namely with respect to $K$-linear functors into categories enriched over $\\mathsf{Vect}_K$.</p>\n" }, { "answer_id": 258745, "score": 4, "text": "<p>I would call these monoidal categories <a href=\"https://ncatlab.org/nlab/show/simple+object\" rel=\"nofollow noreferrer\">simple</a>. But one should exclude the terminal monoidal category, which are <a href=\"http://ncatlab.org/nlab/show/too+simple+to+be+simple\" rel=\"nofollow noreferrer\">too simple to be simple</a>.</p>\n\n<p>Here is a baby example: A commutative monoid, considered as a strict one-object monoidal category (for this to make sense, we need commutativity!), is simple iff it has exactly two regular quotient monoids, the trivial monoid and the monoid itself. Equivalenty, it has exactly two congruence relations.</p>\n\n<p>The only simple commutative groups are $C_p$ for primes $p$. The only finite simple (commutative) monoid, which is not a group, is the multiplicative monoid $(\\{0,1\\},\\cdot,1)$; see Proposition 2.1 in the paper <a href=\"http://www.sciencedirect.com/science/article/pii/0022404989901370\" rel=\"nofollow noreferrer\">The kernel of monoid morphisms</a> by Rhodes and Tilson. Unfortunately, I could not find more about simple commutative monoids.</p>\n" }, { "answer_id": 258644, "score": 4, "text": "<p>Here are some additional thoughts and examples.</p>\n\n<p><strong>1)</strong> The argument presented in the OP applies whenever one has a monoidal category $(\\mathsf{C},\\otimes,1)$ in which the unit object $1$ is a <a href=\"https://ncatlab.org/nlab/show/separator\" rel=\"nofollow noreferrer\">separator</a> and coseparator, and the monoid of scalars $\\mathsf{C}(1,1)$ is itself quotient-free. So under these assumptions, also $\\mathsf{C}$ is quotient-free as a monoidal category.</p>\n\n<p><strong>2)</strong> A similar argument shows that for any field $K\\neq\\mathbb{Z}/2$, the category $\\mathsf{Vect}_K$ has exactly one nontrivial monoidal quotient, namely a category that I will denote by $\\mathsf{Vect}_K/K^\\times$. This quotient is given by identifying any two parallel morphisms that are scalar multiples of each other. (For $K=\\mathbb{Z}/2$, this quotient coincides with the original category, and already $\\mathsf{Vect}_{\\mathbb{Z}/2}$ itself is quotient-free by the same token.)</p>\n\n<p>To prove this claim, note again that the monoidal unit $K\\in\\mathsf{Vect}_K$ is a separator and coseparator, so that any nontrivial quotient of $\\mathsf{Vect}_K$ must also identify two different scalars. But the scalars are the multiplicative monoid $(K,\\cdot)$, which consists of the group of units $(K^\\times,\\cdot)$ together with an additional absorbing element $0\\in K$. Its only nontrivial quotient is the one where all units are identified, with quotient $(\\mathbb{Z}/2,\\cdot)$, which is accidentally also the monoid of scalars of $\\mathsf{Rel}$. By strong monoidality, our original quotient of $\\mathsf{Vect}_K$ must therefore also be a quotient of $\\mathsf{Vect}_K/K^\\times$.</p>\n\n<p>Finally, it remains to be shown that $\\mathsf{Vect}_K/K^\\times$ is quotient-free. So suppose that we have a quotient that identifies two parallel morphisms $[f],[g]:V\\to W$, where I'm working with representatives $f,g\\in\\mathsf{Vect}_K(V,W)$. By suitably pre- and postcomposing with a subspace inclusion and quotient space projection, we may assume that $f=\\mathrm{id}_V$, while $g$ is not a scalar multiple of the identity. In particular, we can choose a basis in which $g$ is not diagonal, resulting in an element $x:K\\to V$ and a functional $t:V\\to K$ such that $tx = tfx =0$, but $tgx\\neq 0$. Hence the quotient must identify the two remaining scalars $0,1\\in\\mathsf{Vect}_K/K^\\times(K,K)$, and we are again done.</p>\n\n<p><strong>3)</strong> The same statement and proof apply to the category $\\mathsf{FVect}_K$ of finite-dimensional vector spaces. But this category is also the free additive $K$-linear category, which makes it quotient-free in a different sense, namely with respect to $K$-linear functors into categories enriched over $\\mathsf{Vect}_K$.</p>\n" }, { "answer_id": 258745, "score": 4, "text": "<p>I would call these monoidal categories <a href=\"https://ncatlab.org/nlab/show/simple+object\" rel=\"nofollow noreferrer\">simple</a>. But one should exclude the terminal monoidal category, which are <a href=\"http://ncatlab.org/nlab/show/too+simple+to+be+simple\" rel=\"nofollow noreferrer\">too simple to be simple</a>.</p>\n\n<p>Here is a baby example: A commutative monoid, considered as a strict one-object monoidal category (for this to make sense, we need commutativity!), is simple iff it has exactly two regular quotient monoids, the trivial monoid and the monoid itself. Equivalenty, it has exactly two congruence relations.</p>\n\n<p>The only simple commutative groups are $C_p$ for primes $p$. The only finite simple (commutative) monoid, which is not a group, is the multiplicative monoid $(\\{0,1\\},\\cdot,1)$; see Proposition 2.1 in the paper <a href=\"http://www.sciencedirect.com/science/article/pii/0022404989901370\" rel=\"nofollow noreferrer\">The kernel of monoid morphisms</a> by Rhodes and Tilson. Unfortunately, I could not find more about simple commutative monoids.</p>\n" } ]
Question: Quotient-free monoidal categories Answer 1: <p>Here are some additional thoughts and examples.</p> <p><strong>1)</strong> The argument presented in the OP applies whenever one has a monoidal category $(\mathsf{C},\otimes,1)$ in which the unit object $1$ is a <a href="https://ncatlab.org/nlab/show/separator" rel="nofollow noreferrer">separator</a> and coseparator, and the monoid of scalars $\mathsf{C}(1,1)$ is itself quotient-free. So under these assumptions, also $\mathsf{C}$ is quotient-free as a monoidal category.</p> <p><strong>2)</strong> A similar argument shows that for any field $K\neq\mathbb{Z}/2$, the category $\mathsf{Vect}_K$ has exactly one nontrivial monoidal quotient, namely a category that I will denote by $\mathsf{Vect}_K/K^\times$. This quotient is given by identifying any two parallel morphisms that are scalar multiples of each other. (For $K=\mathbb{Z}/2$, this quotient coincides with the original category, and already $\mathsf{Vect}_{\mathbb{Z}/2}$ itself is quotient-free by the same token.)</p> <p>To prove this claim, note again that the monoidal unit $K\in\mathsf{Vect}_K$ is a separator and coseparator, so that any nontrivial quotient of $\mathsf{Vect}_K$ must also identify two different scalars. But the scalars are the multiplicative monoid $(K,\cdot)$, which consists of the group of units $(K^\times,\cdot)$ together with an additional absorbing element $0\in K$. Its only nontrivial quotient is the one where all units are identified, with quotient $(\mathbb{Z}/2,\cdot)$, which is accidentally also the monoid of scalars of $\mathsf{Rel}$. By strong monoidality, our original quotient of $\mathsf{Vect}_K$ must therefore also be a quotient of $\mathsf{Vect}_K/K^\times$.</p> <p>Finally, it remains to be shown that $\mathsf{Vect}_K/K^\times$ is quotient-free. So suppose that we have a quotient that identifies two parallel morphisms $[f],[g]:V\to W$, where I'm working with representatives $f,g\in\mathsf{Vect}_K(V,W)$. By suitably pre- and postcomposing with a subspace inclusion and quotient space projection, we may assume that $f=\mathrm{id}_V$, while $g$ is not a scalar multiple of the identity. In particular, we can choose a basis in which $g$ is not diagonal, resulting in an element $x:K\to V$ and a functional $t:V\to K$ such that $tx = tfx =0$, but $tgx\neq 0$. Hence the quotient must identify the two remaining scalars $0,1\in\mathsf{Vect}_K/K^\times(K,K)$, and we are again done.</p> <p><strong>3)</strong> The same statement and proof apply to the category $\mathsf{FVect}_K$ of finite-dimensional vector spaces. But this category is also the free additive $K$-linear category, which makes it quotient-free in a different sense, namely with respect to $K$-linear functors into categories enriched over $\mathsf{Vect}_K$.</p> Answer 2: <p>I would call these monoidal categories <a href="https://ncatlab.org/nlab/show/simple+object" rel="nofollow noreferrer">simple</a>. But one should exclude the terminal monoidal category, which are <a href="http://ncatlab.org/nlab/show/too+simple+to+be+simple" rel="nofollow noreferrer">too simple to be simple</a>.</p> <p>Here is a baby example: A commutative monoid, considered as a strict one-object monoidal category (for this to make sense, we need commutativity!), is simple iff it has exactly two regular quotient monoids, the trivial monoid and the monoid itself. Equivalenty, it has exactly two congruence relations.</p> <p>The only simple commutative groups are $C_p$ for primes $p$. The only finite simple (commutative) monoid, which is not a group, is the multiplicative monoid $(\{0,1\},\cdot,1)$; see Proposition 2.1 in the paper <a href="http://www.sciencedirect.com/science/article/pii/0022404989901370" rel="nofollow noreferrer">The kernel of monoid morphisms</a> by Rhodes and Tilson. Unfortunately, I could not find more about simple commutative monoids.</p> Answer 3: <p>Here are some additional thoughts and examples.</p> <p><strong>1)</strong> The argument presented in the OP applies whenever one has a monoidal category $(\mathsf{C},\otimes,1)$ in which the unit object $1$ is a <a href="https://ncatlab.org/nlab/show/separator" rel="nofollow noreferrer">separator</a> and coseparator, and the monoid of scalars $\mathsf{C}(1,1)$ is itself quotient-free. So under these assumptions, also $\mathsf{C}$ is quotient-free as a monoidal category.</p> <p><strong>2)</strong> A similar argument shows that for any field $K\neq\mathbb{Z}/2$, the category $\mathsf{Vect}_K$ has exactly one nontrivial monoidal quotient, namely a category that I will denote by $\mathsf{Vect}_K/K^\times$. This quotient is given by identifying any two parallel morphisms that are scalar multiples of each other. (For $K=\mathbb{Z}/2$, this quotient coincides with the original category, and already $\mathsf{Vect}_{\mathbb{Z}/2}$ itself is quotient-free by the same token.)</p> <p>To prove this claim, note again that the monoidal unit $K\in\mathsf{Vect}_K$ is a separator and coseparator, so that any nontrivial quotient of $\mathsf{Vect}_K$ must also identify two different scalars. But the scalars are the multiplicative monoid $(K,\cdot)$, which consists of the group of units $(K^\times,\cdot)$ together with an additional absorbing element $0\in K$. Its only nontrivial quotient is the one where all units are identified, with quotient $(\mathbb{Z}/2,\cdot)$, which is accidentally also the monoid of scalars of $\mathsf{Rel}$. By strong monoidality, our original quotient of $\mathsf{Vect}_K$ must therefore also be a quotient of $\mathsf{Vect}_K/K^\times$.</p> <p>Finally, it remains to be shown that $\mathsf{Vect}_K/K^\times$ is quotient-free. So suppose that we have a quotient that identifies two parallel morphisms $[f],[g]:V\to W$, where I'm working with representatives $f,g\in\mathsf{Vect}_K(V,W)$. By suitably pre- and postcomposing with a subspace inclusion and quotient space projection, we may assume that $f=\mathrm{id}_V$, while $g$ is not a scalar multiple of the identity. In particular, we can choose a basis in which $g$ is not diagonal, resulting in an element $x:K\to V$ and a functional $t:V\to K$ such that $tx = tfx =0$, but $tgx\neq 0$. Hence the quotient must identify the two remaining scalars $0,1\in\mathsf{Vect}_K/K^\times(K,K)$, and we are again done.</p> <p><strong>3)</strong> The same statement and proof apply to the category $\mathsf{FVect}_K$ of finite-dimensional vector spaces. But this category is also the free additive $K$-linear category, which makes it quotient-free in a different sense, namely with respect to $K$-linear functors into categories enriched over $\mathsf{Vect}_K$.</p> Answer 4: <p>I would call these monoidal categories <a href="https://ncatlab.org/nlab/show/simple+object" rel="nofollow noreferrer">simple</a>. But one should exclude the terminal monoidal category, which are <a href="http://ncatlab.org/nlab/show/too+simple+to+be+simple" rel="nofollow noreferrer">too simple to be simple</a>.</p> <p>Here is a baby example: A commutative monoid, considered as a strict one-object monoidal category (for this to make sense, we need commutativity!), is simple iff it has exactly two regular quotient monoids, the trivial monoid and the monoid itself. Equivalenty, it has exactly two congruence relations.</p> <p>The only simple commutative groups are $C_p$ for primes $p$. The only finite simple (commutative) monoid, which is not a group, is the multiplicative monoid $(\{0,1\},\cdot,1)$; see Proposition 2.1 in the paper <a href="http://www.sciencedirect.com/science/article/pii/0022404989901370" rel="nofollow noreferrer">The kernel of monoid morphisms</a> by Rhodes and Tilson. Unfortunately, I could not find more about simple commutative monoids.</p>
What mode of convergence is this?
258,871
8
https://mathoverflow.net/questions/258871/what-mode-of-convergence-is-this
<p>I'm interested in a new (to me) mode of convergence which is stronger than convergence in measure/probability. I want to know if it has a name and if it is used much in the literature. I will write my question twice using both probability theory notation and measure theory notation to ensure that I reach the correct audience.</p> <hr /> <p><strong>Question in measure theory notation</strong></p> <p>Let <span class="math-container">$f_n:\mathbb{X} \to \mathbb{Y}$</span> be a sequence of measurable functions where <span class="math-container">$(\mathbb{X},d_\mathbb{X})$</span> and <span class="math-container">$(\mathbb{Y},d_\mathbb{Y})$</span> are complete separable metric spaces (if it helps, assume <span class="math-container">$\mathbb{X}=\mathbb{Y}=\mathbb{R}$</span>). Let <span class="math-container">$\mu$</span> be a probability measure on <span class="math-container">$\mathbb{X}$</span>.</p> <p>It is well-known that <span class="math-container">$f_n$</span> <em>converges in measure</em> to a measurable function <span class="math-container">$f_n:\mathbb{X} \to \mathbb{Y}$</span> if <span class="math-container">$\mu \{x:d_\mathbb{Y}(f_n, f) &gt; \varepsilon\} \to 0$</span> as <span class="math-container">$n \to \infty$</span> for all <span class="math-container">$\varepsilon &gt; 0$</span>.</p> <p>Recall, that every function <span class="math-container">$f$</span> induces a <span class="math-container">$\sigma$</span>-algebra <span class="math-container">$$\sigma(f) = \{A \subseteq \mathbb{X} | A = f^{-1}(B), B\subseteq \mathbb{Y} \text{ measurable}\}.$$</span> My new mode of convergence will require that <span class="math-container">$f_n \to f$</span> in measure and that (informally) <span class="math-container">$\sigma(f_n) \to \sigma(f)$</span>. This latter condition can be made formal using conditional expectation as follows.</p> <p>Recall that the conditional expectation of a measurable function <span class="math-container">$g:\mathbb{X} \to \mathbb{R}$</span> conditioned on the sigma-algebra <span class="math-container">$\sigma(f)$</span>, written <span class="math-container">$E(g|\sigma(f))$</span> or just <span class="math-container">$E(g|f)$</span> is the (<span class="math-container">$\mu$</span>-a.e.) unique <span class="math-container">$\sigma(f)$</span>-measurable function <span class="math-container">$h$</span> given by <span class="math-container">$$\int_A h\, d\mu = \int_A f d\mu \qquad \text{for $A \in \sigma(f)$}.$$</span></p> <blockquote> <p><strong>Definition.</strong> Say <span class="math-container">$f_n$</span> <em>converges in ???</em> to a measurable function <span class="math-container">$f_n:\mathbb{X} \to \mathbb{Y}$</span> if <span class="math-container">$f_n$</span> converges to <span class="math-container">$f$</span> in measure and for all <span class="math-container">$\mu$</span>-integrable functions <span class="math-container">$g:\mathbb{X} \to \mathbb{R}$</span>, we have that <span class="math-container">$E(g | f_n)$</span> converges to <span class="math-container">$E(g | f)$</span> in measure.</p> <p><strong>Question.</strong> Does &quot;converges in ???&quot; have a formal name (for the mode of convergence or for the topology it induces on the space of measurable functions modulo a.e.-equivalence)? Is there any places or standard resources where it is used?</p> </blockquote> <hr /> <p><strong>Question in probability notation</strong></p> <p>Let <span class="math-container">$X_n$</span> be a sequence of random variables taking values in a complete separable metric space <span class="math-container">$(\mathbb{X},d)$</span> (if it helps, assume <span class="math-container">$\mathbb{X}=\mathbb{R}$</span>).</p> <p>It is well-known that <span class="math-container">$X_n$</span> <em>converges in probability</em> to a random variable <span class="math-container">$X$</span> if <span class="math-container">$P\{d(X_n, X) &gt; \varepsilon\} \to 0$</span> as <span class="math-container">$n \to \infty$</span> for all <span class="math-container">$\varepsilon &gt; 0$</span>.</p> <blockquote> <p><strong>Definition.</strong> Say <span class="math-container">$X_n$</span> <em>converges in ???</em> to a random variable <span class="math-container">$X$</span> if <span class="math-container">$X_n$</span> converges to <span class="math-container">$X$</span> in measure and for all bounded continuous functions <span class="math-container">$g:\mathbb{X}^\infty \to \mathbb{R}$</span>, we have that <span class="math-container">$E(g(X,X_1,X_2,\ldots) | X_n)$</span> converges to <span class="math-container">$E(g(X,X_1,X_2,\ldots) | X)$</span> in probability.</p> <p><strong>Question.</strong> Does &quot;converges in ???&quot; have a formal name (for the mode of convergence or for the topology it induces on the space of random variables modulo a.s.-equivalence)? Is there any places or standard resources where it is used?</p> </blockquote> <hr /> <p><strong>Bonus Question</strong></p> <p>As I hinted to above, this approach also gives a notion of convergence of sigma-algebras (for a particular measure space <span class="math-container">$(\mathbb{X},\mu)$</span> as above). Fix a sequence of sub-sigma algebras <span class="math-container">$\mathcal{F}_n$</span> of the Borel sigma-algebra. Say that <em><span class="math-container">$\mathcal{F}_n$</span> converges to <span class="math-container">$\mathcal{F}$</span></em> if <span class="math-container">$E(g | \mathcal{F}_n)$</span> converges to <span class="math-container">$E(g | \mathcal{F})$</span> in measure for all <span class="math-container">$\mu$</span>-integrable functions <span class="math-container">$g:\mathbb{X}\to\mathbb{R}$</span>. (For example, by the Levy 0-1 law, if <span class="math-container">$\mathcal{F}_n$</span> is a filtration, then <span class="math-container">$\mathcal{F}_n$</span> converges to <span class="math-container">$\mathcal{F}_\infty$</span> in this sense.)</p> <blockquote> <p><strong>Question.</strong> Does this notion of convergence on sigma-algebras have a formal name (for the mode of convergence or for the topology it induces on the space of sigma-algebras modulo a.e.-equivalence)? Is there any places or standard resources where it is used?</p> </blockquote> <hr /> <p><strong>Appendix 1</strong></p> <p>This mode of convergence is not the same as convergence in measure/probability. Here is a simple example. Let <span class="math-container">$U$</span> be a random variable uniformly distributed in <span class="math-container">$[0,1]$</span>. Let <span class="math-container">$X_n = U/n$</span> and <span class="math-container">$X=0$</span>. It is clear that <span class="math-container">$X_n$</span> converges to <span class="math-container">$X$</span> in probability, but it is not true that <span class="math-container">$X_n$</span> converges to <span class="math-container">$X$</span> in this new mode of convergence, since <span class="math-container">$E(X_1|X_n) = E(U|X_n) = U$</span>, but <span class="math-container">$E(X_1|X) = E(U|X) = E(U) = 1/2$</span>.</p> <hr /> <p><strong>Appendix 2</strong></p> <p>It should be pointed out that the measure theory and probability theory definitions I gave differ in more than just notation. For one, the probability theory definition doesn't use the space <span class="math-container">$(\Omega,P)$</span> at all. However, if <span class="math-container">$\Omega$</span> is a completely separable measure space, <span class="math-container">$P$</span> is a Borel probability measure, and <span class="math-container">$X$</span> and <span class="math-container">$X_n$</span> are all Borel measurable, then the two definitions are equivalent. Here is a proof sketch.</p> <p>Assume <span class="math-container">$E(h|X_n)$</span> converges in probability/measure to <span class="math-container">$E(h|X)$</span> for all <span class="math-container">$P$</span>-integrable continuous <span class="math-container">$h:\Omega \to \mathbb{R}$</span>. Then <span class="math-container">$g(X,X_1,X_2,\ldots)$</span> is <span class="math-container">$P$</span>-integrable and therefore, <span class="math-container">$E(g(X,X_1,X_2,\ldots) | X_n)$</span> converges to <span class="math-container">$E(g(X,X_1,X_2,\ldots) | X)$</span> in probability.</p> <p>Conversely assume <span class="math-container">$E(g(X,X_1,X_2,\ldots) | X_n)$</span> converges to <span class="math-container">$E(g(X,X_1,X_2,\ldots) | X)$</span> for all bounded continuous functions <span class="math-container">$g$</span>. Let <span class="math-container">$f$</span> be a <span class="math-container">$P$</span>-integrable function. Then using the tower property of conditional expectation we have that <span class="math-container">$E(f|X_n) = E(E(f|X,X_1,X_2 \ldots)|X_n)$</span>. The function <span class="math-container">$E(f|X,X_1,X_2 \ldots)$</span> is <span class="math-container">$P$</span>-integrable and therefore, there is a <span class="math-container">$P_(X,X_1,X_2,\ldots)$</span>-integrable function <span class="math-container">$h:\mathbb{X}^\infty \to \mathbb{R}$</span> such that <span class="math-container">$E(f|X,X_1,X_2 \ldots) = h(X,X_1,X_2,\ldots)$</span> a.s.</p> <p>So it remains to show that <span class="math-container">$E(h(X,X_1,X_2,\ldots) | X_n)$</span> converges to <span class="math-container">$E(h(X,X_1,X_2,\ldots) | X)$</span> in probability. The main idea is that since <span class="math-container">$h$</span> is integrable, we can approximate it with a continuous functions <span class="math-container">$g$</span> which are close in the <span class="math-container">$L_1$</span>-norm. (The formal computation is left to the reader.)</p>
[ { "answer_id": 259018, "score": 3, "text": "<p>I'm not sure if this answers the bonus question, but topologies like that have been studied in mathematical economics to model private information of agents. Chapter 9 of the \"Handbook of Applied Analysis\" by Nikolaos S. Papageorgiou and Sophia Th. Kyritsi-Yiallourou contains a lot of material on the subject and you may find there what you need, it certainly studies the variant of your topology in which convergence in measure is replaced by $L_1$-convergence.</p>\n" }, { "answer_id": 259633, "score": 0, "text": "<p>I don't believe it is standard in probability theory to denote that mode of convergence in a specific manner. That being said, in the special case where the conditioning $\\sigma$-fields are indexed and filtered by any set $I\\subset \\mathbb{R}$, there is convergence of $E(g|\\mathcal{F_t})$, $t\\in I$ along any increasing or decreasing sequence. In this case, one talks about $\\textit{martingale convergence}$, or $\\textit{reverse-martingale convergence}$, respectively. </p>\n" }, { "answer_id": 259018, "score": 3, "text": "<p>I'm not sure if this answers the bonus question, but topologies like that have been studied in mathematical economics to model private information of agents. Chapter 9 of the \"Handbook of Applied Analysis\" by Nikolaos S. Papageorgiou and Sophia Th. Kyritsi-Yiallourou contains a lot of material on the subject and you may find there what you need, it certainly studies the variant of your topology in which convergence in measure is replaced by $L_1$-convergence.</p>\n" }, { "answer_id": 259633, "score": 0, "text": "<p>I don't believe it is standard in probability theory to denote that mode of convergence in a specific manner. That being said, in the special case where the conditioning $\\sigma$-fields are indexed and filtered by any set $I\\subset \\mathbb{R}$, there is convergence of $E(g|\\mathcal{F_t})$, $t\\in I$ along any increasing or decreasing sequence. In this case, one talks about $\\textit{martingale convergence}$, or $\\textit{reverse-martingale convergence}$, respectively. </p>\n" } ]
Question: What mode of convergence is this? Answer 1: <p>I'm not sure if this answers the bonus question, but topologies like that have been studied in mathematical economics to model private information of agents. Chapter 9 of the "Handbook of Applied Analysis" by Nikolaos S. Papageorgiou and Sophia Th. Kyritsi-Yiallourou contains a lot of material on the subject and you may find there what you need, it certainly studies the variant of your topology in which convergence in measure is replaced by $L_1$-convergence.</p> Answer 2: <p>I don't believe it is standard in probability theory to denote that mode of convergence in a specific manner. That being said, in the special case where the conditioning $\sigma$-fields are indexed and filtered by any set $I\subset \mathbb{R}$, there is convergence of $E(g|\mathcal{F_t})$, $t\in I$ along any increasing or decreasing sequence. In this case, one talks about $\textit{martingale convergence}$, or $\textit{reverse-martingale convergence}$, respectively. </p> Answer 3: <p>I'm not sure if this answers the bonus question, but topologies like that have been studied in mathematical economics to model private information of agents. Chapter 9 of the "Handbook of Applied Analysis" by Nikolaos S. Papageorgiou and Sophia Th. Kyritsi-Yiallourou contains a lot of material on the subject and you may find there what you need, it certainly studies the variant of your topology in which convergence in measure is replaced by $L_1$-convergence.</p> Answer 4: <p>I don't believe it is standard in probability theory to denote that mode of convergence in a specific manner. That being said, in the special case where the conditioning $\sigma$-fields are indexed and filtered by any set $I\subset \mathbb{R}$, there is convergence of $E(g|\mathcal{F_t})$, $t\in I$ along any increasing or decreasing sequence. In this case, one talks about $\textit{martingale convergence}$, or $\textit{reverse-martingale convergence}$, respectively. </p>
What is the mathematical name for Haskell&#39;s Alternative Functor
259,120
8
https://mathoverflow.net/questions/259120/what-is-the-mathematical-name-for-haskells-alternative-functor
<p>Haskell's Applicative Functor is called Lax Monoidal Functor in mathematics.</p> <p>What is Haskell's Alternative Functor called in mathematics?</p> <p>Recall that Haskell's Alternative Functor is defined as follows:</p> <pre><code>class Applicative f =&gt; Alternative f where -- | The identity of '&lt;|&gt;' empty :: f a -- | An associative binary operation (&lt;|&gt;) :: f a -&gt; f a -&gt; f a </code></pre> <p>Edit: In other words, an Alternative Functor <code>f</code> is an Applicative Functor such that for every type <code>a</code>, <code>f a</code> is a monoid, where <code>&lt;|&gt;</code> is the binary operation of type <code>f a -&gt; f a -&gt; f a</code> and <code>empty</code> is the identity of type <code>f a</code>. Uncurrying the type of <code>&lt;|&gt;</code>, we get $$f\ a \times f\ a \rightarrow f\ a$$ and <code>empty</code> can be equivalently defined to have type $$1 \rightarrow f\ a$$ In the Haskell definition, the type variable <code>a</code> is universally quantified. As far as I know, this translates to the requirement that they are natural in <code>a</code>.</p>
[ { "answer_id": 259163, "score": 9, "text": "<p>This is not a complete answer, just a translation of Haskell to math so that the resident category theorists can tell what the question is.</p>\n\n<p>As is usual in such cases, the terminology and the concepts in Haskell are slightly modified concepts from mathematics. It is best in the first iteration to look just at the broad similarities and not worry so much about the details.</p>\n\n<p>Recall from <a href=\"http://www.staff.city.ac.uk/~ross/papers/Applicative.pdf\" rel=\"noreferrer\">Section 7 of the paper</a> by Connor McBride and Ross Paterson that Haskell's <code>Applicative</code> is a lax monoidal functor (kind of as the Haskell definition uses internal Hom-sets where one would expect the external ones).</p>\n\n<p>The definition of <code>Alternative</code> translates into: a functor $F : \\mathcal{C} \\to \\mathcal{C}$, where $C$ is at least cartesian, together with natural transformations $e : 1 \\to F$ and $m : F({-}) \\times F({-}) \\to F$, or concretely, for every object $A \\in \\mathcal{C}$ we have arrows\n$$e_A : 1 \\to F(A)$$\nand\n$$m_A : F(A) \\times F(A) \\to F(A)$$\nsuch that $(F(A), e_A, m_A)$ is a monoid for all $A$, naturally in $A$. What would you call such a thing?</p>\n\n<p>The next question to ask is whether there is anything extra about having a lax monoidal functor which also has the above structure of monoids? I don't know off the top of my head. If there is another view of the same situation, it will surely be useful for various Haskell hacking tricks. Haskell people are very good at using category-theoretic algebra for all sorts of cool purposes.</p>\n\n<p>Let me also explain on the difference between parametricity and naturality. When we define a \"functor\" $F$ in Haskell, that is not really a functor. It is acertain mapping from types to types which acts on internal hom-sets as a functor would. In addition, this mapping has a strong property known as <em><a href=\"https://en.wikipedia.org/wiki/Parametricity\" rel=\"noreferrer\">parametricity</a></em> which says that $F$ behaves \"in the same way\" on all types. Parametricity is quite similar to naturality, and there is a way to view it semantically as such, but it is not precisely naturality. This is not a detail into which we should worry about here, though.</p>\n" }, { "answer_id": 343993, "score": 1, "text": "<p>A slightly different set of primitive operations that we can look at is the following:</p>\n\n\n\n<pre><code>class Functor f =&gt; Alternative' f\n where\n unit :: () -&gt; f Void\n zip :: (f a, f b) -&gt; f (Either a b)\n</code></pre>\n\n<p>This represents a lax monoidal functor from the monoidal category of <code>-&gt;</code> under the tensor <code>Either</code> to the monoidal category of <code>-&gt;</code> under the tensor <code>(,)</code>. The laws are the usual laws for the coherence maps of lax monoidal functors, including those that demand conformance with the functor (i.e. you can crib them with minor adjustments from the <code>Applicative</code> typeclass).</p>\n\n<p>By contrast, the operations and laws of the <code>Alternative</code> typeclass of today seem to amount to <code>type Alternative f = Monoid (forall a. f a)</code>. We can witness a pair of maps that convert between the primitive operations of <code>Alternative'</code> and <code>Alternative</code>:</p>\n\n\n\n<pre><code>instance Alternative f =&gt; Alternative' f\n where\n unit _ = empty\n zip (fa, fb) = (Left &lt;<span class=\"math-container\">$&gt; fa) &lt;|&gt; (Right &lt;$</span>&gt; fb)\n\ninstance Alternative' f =&gt; Alternative f\n where\n empty = absurd &lt;<span class=\"math-container\">$&gt; unit ()\n fa1 &lt;|&gt; fa2 = either id id &lt;$</span>&gt; zip (fa1, fa2)\n</code></pre>\n\n<p>However, it seems likely that this is not truly an isomorphism, due to the difference in constraints and laws between the two typeclasses.</p>\n\n<p>On one hand, <code>Alternative</code> is more permissive than <code>Alternative'</code>, since it does not demand any form of compatibility between the monoid and the functor, whereas the commutative diagrams of a lax monoidal functor do.</p>\n\n<p>On, the other hand, <code>Alternative</code> demands <code>Applicative</code>, a constraint that <code>Alternative'</code> does not demand. When viewed through the lens of lax monoidal functors, <code>Applicative f</code> simply means that <code>f</code> coheres the <code>(,)</code> monoidal structure to <code>(,)</code>, and <code>Alternative f</code> means that it coheres the <code>Either</code> monoidal structure to <code>(,)</code>. These are orthogonal claims, so there is no need for <code>Alternative' f</code> to demand a <code>Applicative f</code> constraint.</p>\n\n<p>That said, I conjecture that most types that support <code>Alternative</code> instances also support a lawful instance of <code>Alternative'</code>, and so you can imagine them to be \"applicatives for different tensor\".</p>\n" }, { "answer_id": 351030, "score": 0, "text": "<p>This might not be the right approach, but here's my take on the question:</p>\n\n<p>If <span class=\"math-container\">$(\\mathcal{C}, \\otimes, I, \\dots)$</span> is a (strict, for the sake of readability) monoidal category, we can endow <span class=\"math-container\">$[\\mathcal{C}, \\mathcal{C}]$</span> with two different monoidal structures:</p>\n\n<ol>\n<li><p>The \"usual\" monoidal structure, where <span class=\"math-container\">$\\circ$</span> is the product and <span class=\"math-container\">$\\mathbf{1}_\\mathcal{C}$</span> is the unit. This is the monoidal category whose monoids are called monads. Notice how this doesn't actually use the fact that <span class=\"math-container\">$\\mathcal{C}$</span> is endowed with a monoidal structure.</p></li>\n<li><p>The \"other\" (I sometimes refer to it as \"pointwise\") monoidal structure has \"pointwise <span class=\"math-container\">$\\otimes$</span>\" (let's call it <span class=\"math-container\">$\\boxtimes$</span>) as product, and <span class=\"math-container\">$\\Delta_I$</span> (the constant functor on the unit object of <span class=\"math-container\">$\\mathcal{C}$</span>) as monoidal units. In this category, monoids are related to what Haskell programmers call \"alternative\" functors, but without the part about being \"applicative\" and with one additional property. This can be further generalized to the case of functors <span class=\"math-container\">$\\mathcal{B} \\to \\mathcal{C}$</span> where <span class=\"math-container\">$\\mathcal{B}$</span> is <em>not</em> required to be monoidal, and everything that follows should work just fine (but I won't venture there).</p></li>\n</ol>\n\n<p>Let me elaborate on the latter, since the former is (probably) already well-known to most haskell programmers here (other than being significantly less relevant):</p>\n\n<p>First we'll need to properly define <span class=\"math-container\">$\\boxtimes$</span>: if we fix functors <span class=\"math-container\">$\\mathcal{F}, \\mathcal{G} : \\mathcal{C} \\to \\mathcal{C}$</span> we can define <span class=\"math-container\">$(\\mathcal{F} \\boxtimes \\mathcal{G}) (a) := \\mathcal{F}(a) \\otimes \\mathcal{G}(a)$</span> (the definition for morphisms is analogous). This, together with <span class=\"math-container\">$\\Delta_I$</span> as a unit, defines a monoidal category: all axioms (and further properties, such as symmetry and strictness) follow trivially from the monoidal structure on <span class=\"math-container\">$\\mathcal{C}$</span>. </p>\n\n<p>Now, monoids in this monoidal category are what haskell programmers call \"alternative\" functors, but that's not all: they are so in a functorial way, meaning that if <span class=\"math-container\">$f : a \\to b : \\mathcal{C}$</span> and <span class=\"math-container\">$\\mathcal{M}$</span> is a monoid in <span class=\"math-container\">$([\\mathcal{C}, \\mathcal{C}], \\boxtimes, \\Delta_I)$</span> then <span class=\"math-container\">$\\mathcal{M}(f)$</span> is a morphism of monoids. To see this, note how <a href=\"https://tikzcd.yichuanshen.de/#N4Igdg9gJgpgziAXAbVABwnAlgFyxMJZABgBpiBdUkANwEMAbAVxiRAB12BbOnACwDGjYAFkAvgAJOAIwgAPPF3hT2AERgMcdAPoBJEGNLpMufIRQBGclVqMWbTj35CGosQaMgM2PASIAma2p6ZlZEDjUNLT0VWQUsJTgVJ0FhcQ9jHzMiMgsbEPtwx15U13FY+UVlYuc090NM0z9LUjzguzCIlJc3DK8TX3NkQLbbUIduEp7ymUqE6snasvqbGCgAc3giUAAzACcILiQyEBwIJCsxwoi9vnPqBjppDQAFAezwvax1vhw+-cOSECp3OiEuBU6nEeXGkUDoIAeT1e72aIAYMB2fwaIABR0QwLOSAArO1xkV2DA0NgGAQKvFEskStIdsALGJtDVSr1Ec8GG8sqivj8sZ5ccdqITEABmHnIgXmNEYv6k66c5ms9mc6aSWb0haU6l+bFi6US0EAFllfJRCvRmIRV0h3CY-wOeJJIKQlrRSOt8rYQt+DohEy4LuNbouZq9KqdKXVbI5iy56TEFDEQA\" rel=\"nofollow noreferrer\">this diagram</a> needs to commute in a pointwise fashion, and the same applies to <a href=\"https://tikzcd.yichuanshen.de/#N4Igdg9gJgpgziAXAbVABwnAlgFyxMJZABgBpiBdUkANwEMAbAVxiRAB12BbOnACwDGjYAFkAvgAJOAIwgAPPF3hTuvQcPErZCrErgqe-IQ1FiQY0uky58hFGQCMVWoxZtOh9Sc0z5i5R5qxqbmliAY2HgERABM5M70zKyIHKpGGpK+OnoGQRmhVpG2saRO1IluKYHp3mZizjBQAObwRKAAZgBOEFxIZCA4EEgO5a7JqVxMWn66AWl80u3ADmIA+tVeIRYd3b2IcQNDiADM1Ax00jAMAArWUXYgnVhNfDggo0nu3EwFIF09w2ogyQpxcnyq31+-z2-WB+zOFyutyK0RSTxebw+lQmakWyzWG2CPnY2n8+g8P3qYiAA\" rel=\"nofollow noreferrer\">this other diagram</a> too. We're in the strict case, but nothing significant changes in the general one (except that you need to keep track of the coherence laws).</p>\n\n<p>A more abstract (but equivalent) formulation is that each object <span class=\"math-container\">$\\mathcal{M} : \\mathbf{Mon}([\\mathcal{C}, \\mathcal{C}])$</span> naturally defines a functor <span class=\"math-container\">$\\hat{\\mathcal{M}} : \\mathcal{C} \\to \\mathbf{Mon}(\\mathcal{C})$</span>, and this point of view leads to the observation that this actually defines a functor <span class=\"math-container\">$(\\hat{-}) : \\mathbf{Mon}([\\mathcal{C}, \\mathcal{C}]) \\to [\\mathcal{C}, \\mathbf{Mon}(\\mathcal{C})]$</span>. </p>\n\n<p>What this amounts to is just that a \"monoidal\" natural transformation <span class=\"math-container\">$\\nu : \\mathcal{M} \\to \\mathcal{N} : \\mathbf{Mon}([\\mathcal{C}, \\mathcal{C}])$</span> between monoid object in this \"pointwise\" monoidal category of (endo)functors also defines a natural transformation <span class=\"math-container\">$\\hat{\\nu} : \\hat{\\mathcal{M}} \\to \\hat{\\mathcal{N}} : [\\mathcal{C}, \\mathbf{Mon}(\\mathcal{C})]$</span>: the argument is similar but you need to look at the diagrams for <a href=\"https://ncatlab.org/nlab/show/monoid+in+a+monoidal+category#morphism_of_monoids\" rel=\"nofollow noreferrer\">morphisms of monoids</a>.</p>\n\n<p>Things are not over yet: it seems that <span class=\"math-container\">$(\\hat{-})$</span> is actually an equivalence of categories! You can convince yourself of this if you squint hard enough, and the argument is roughly \"brute force\":</p>\n\n<p>Let's construct a strict inverse to <span class=\"math-container\">$(\\hat{-})$</span>, call it <span class=\"math-container\">$(\\tilde{-})$</span>: take <span class=\"math-container\">$\\mathcal{F} : \\mathcal{C} \\to \\mathbf{Mon}(\\mathcal{C})$</span>, compose it with the forgetful functor <span class=\"math-container\">$\\mathbf{U} : \\mathbf{Mon}(\\mathcal{C}) \\to \\mathcal{C}$</span>. We'll now use the \"forgotten structure\" on <span class=\"math-container\">$\\mathcal{F}$</span> to endow <span class=\"math-container\">$\\mathbf{U} \\circ \\mathcal{F}$</span> with the structure of a monoid in <span class=\"math-container\">$[\\mathcal{C}, \\mathcal{C}]$</span>: to do this, notice how the monoid structure on <span class=\"math-container\">$\\mathcal{F}(a)$</span> (for <span class=\"math-container\">$a : \\mathcal{C}$</span>) needs to be such that each morphism <span class=\"math-container\">$\\mathcal{F}(f) : \\mathcal{F}(a) \\to \\mathcal{F}(b)$</span> is a morphism of monoids: this amounts to saying that the correspondences <span class=\"math-container\">$a \\mapsto \\mathbf{m}^{\\mathcal{F}(a)}$</span> and <span class=\"math-container\">$a \\mapsto \\mathbf{e}^{\\mathcal{F}(a)}$</span> are natural in <span class=\"math-container\">$a$</span> and hence natural transformations (its now trivial to check that these endow <span class=\"math-container\">$\\mathbf{U} \\circ \\mathcal{F}$</span> with the structure of a monoid in <span class=\"math-container\">$[\\mathcal{C}, \\mathcal{C}]$</span>).</p>\n\n<p>We now need to check that <span class=\"math-container\">$(\\hat{-}) \\circ (\\tilde{-}) = \\mathbf{1}_{[\\mathcal{C}, \\mathbf{Mon}(\\mathcal{C})]}$</span> and <span class=\"math-container\">$(\\tilde{-} \\circ \\hat{-}) = \\mathbf{1}_{\\mathbf{Mon}([\\mathcal{C}, \\mathcal{C}])}$</span>: they should both be long and trivial computations, and I'll (boldly) omit them.</p>\n\n<p>Edit: this can probably be \"merged\" to <a href=\"https://mathoverflow.net/a/259163/111265\">the other answer by Andrej Bauer</a> to give a more complete answer</p>\n" }, { "answer_id": 259163, "score": 9, "text": "<p>This is not a complete answer, just a translation of Haskell to math so that the resident category theorists can tell what the question is.</p>\n\n<p>As is usual in such cases, the terminology and the concepts in Haskell are slightly modified concepts from mathematics. It is best in the first iteration to look just at the broad similarities and not worry so much about the details.</p>\n\n<p>Recall from <a href=\"http://www.staff.city.ac.uk/~ross/papers/Applicative.pdf\" rel=\"noreferrer\">Section 7 of the paper</a> by Connor McBride and Ross Paterson that Haskell's <code>Applicative</code> is a lax monoidal functor (kind of as the Haskell definition uses internal Hom-sets where one would expect the external ones).</p>\n\n<p>The definition of <code>Alternative</code> translates into: a functor $F : \\mathcal{C} \\to \\mathcal{C}$, where $C$ is at least cartesian, together with natural transformations $e : 1 \\to F$ and $m : F({-}) \\times F({-}) \\to F$, or concretely, for every object $A \\in \\mathcal{C}$ we have arrows\n$$e_A : 1 \\to F(A)$$\nand\n$$m_A : F(A) \\times F(A) \\to F(A)$$\nsuch that $(F(A), e_A, m_A)$ is a monoid for all $A$, naturally in $A$. What would you call such a thing?</p>\n\n<p>The next question to ask is whether there is anything extra about having a lax monoidal functor which also has the above structure of monoids? I don't know off the top of my head. If there is another view of the same situation, it will surely be useful for various Haskell hacking tricks. Haskell people are very good at using category-theoretic algebra for all sorts of cool purposes.</p>\n\n<p>Let me also explain on the difference between parametricity and naturality. When we define a \"functor\" $F$ in Haskell, that is not really a functor. It is acertain mapping from types to types which acts on internal hom-sets as a functor would. In addition, this mapping has a strong property known as <em><a href=\"https://en.wikipedia.org/wiki/Parametricity\" rel=\"noreferrer\">parametricity</a></em> which says that $F$ behaves \"in the same way\" on all types. Parametricity is quite similar to naturality, and there is a way to view it semantically as such, but it is not precisely naturality. This is not a detail into which we should worry about here, though.</p>\n" }, { "answer_id": 343993, "score": 1, "text": "<p>A slightly different set of primitive operations that we can look at is the following:</p>\n\n\n\n<pre><code>class Functor f =&gt; Alternative' f\n where\n unit :: () -&gt; f Void\n zip :: (f a, f b) -&gt; f (Either a b)\n</code></pre>\n\n<p>This represents a lax monoidal functor from the monoidal category of <code>-&gt;</code> under the tensor <code>Either</code> to the monoidal category of <code>-&gt;</code> under the tensor <code>(,)</code>. The laws are the usual laws for the coherence maps of lax monoidal functors, including those that demand conformance with the functor (i.e. you can crib them with minor adjustments from the <code>Applicative</code> typeclass).</p>\n\n<p>By contrast, the operations and laws of the <code>Alternative</code> typeclass of today seem to amount to <code>type Alternative f = Monoid (forall a. f a)</code>. We can witness a pair of maps that convert between the primitive operations of <code>Alternative'</code> and <code>Alternative</code>:</p>\n\n\n\n<pre><code>instance Alternative f =&gt; Alternative' f\n where\n unit _ = empty\n zip (fa, fb) = (Left &lt;<span class=\"math-container\">$&gt; fa) &lt;|&gt; (Right &lt;$</span>&gt; fb)\n\ninstance Alternative' f =&gt; Alternative f\n where\n empty = absurd &lt;<span class=\"math-container\">$&gt; unit ()\n fa1 &lt;|&gt; fa2 = either id id &lt;$</span>&gt; zip (fa1, fa2)\n</code></pre>\n\n<p>However, it seems likely that this is not truly an isomorphism, due to the difference in constraints and laws between the two typeclasses.</p>\n\n<p>On one hand, <code>Alternative</code> is more permissive than <code>Alternative'</code>, since it does not demand any form of compatibility between the monoid and the functor, whereas the commutative diagrams of a lax monoidal functor do.</p>\n\n<p>On, the other hand, <code>Alternative</code> demands <code>Applicative</code>, a constraint that <code>Alternative'</code> does not demand. When viewed through the lens of lax monoidal functors, <code>Applicative f</code> simply means that <code>f</code> coheres the <code>(,)</code> monoidal structure to <code>(,)</code>, and <code>Alternative f</code> means that it coheres the <code>Either</code> monoidal structure to <code>(,)</code>. These are orthogonal claims, so there is no need for <code>Alternative' f</code> to demand a <code>Applicative f</code> constraint.</p>\n\n<p>That said, I conjecture that most types that support <code>Alternative</code> instances also support a lawful instance of <code>Alternative'</code>, and so you can imagine them to be \"applicatives for different tensor\".</p>\n" }, { "answer_id": 351030, "score": 0, "text": "<p>This might not be the right approach, but here's my take on the question:</p>\n\n<p>If <span class=\"math-container\">$(\\mathcal{C}, \\otimes, I, \\dots)$</span> is a (strict, for the sake of readability) monoidal category, we can endow <span class=\"math-container\">$[\\mathcal{C}, \\mathcal{C}]$</span> with two different monoidal structures:</p>\n\n<ol>\n<li><p>The \"usual\" monoidal structure, where <span class=\"math-container\">$\\circ$</span> is the product and <span class=\"math-container\">$\\mathbf{1}_\\mathcal{C}$</span> is the unit. This is the monoidal category whose monoids are called monads. Notice how this doesn't actually use the fact that <span class=\"math-container\">$\\mathcal{C}$</span> is endowed with a monoidal structure.</p></li>\n<li><p>The \"other\" (I sometimes refer to it as \"pointwise\") monoidal structure has \"pointwise <span class=\"math-container\">$\\otimes$</span>\" (let's call it <span class=\"math-container\">$\\boxtimes$</span>) as product, and <span class=\"math-container\">$\\Delta_I$</span> (the constant functor on the unit object of <span class=\"math-container\">$\\mathcal{C}$</span>) as monoidal units. In this category, monoids are related to what Haskell programmers call \"alternative\" functors, but without the part about being \"applicative\" and with one additional property. This can be further generalized to the case of functors <span class=\"math-container\">$\\mathcal{B} \\to \\mathcal{C}$</span> where <span class=\"math-container\">$\\mathcal{B}$</span> is <em>not</em> required to be monoidal, and everything that follows should work just fine (but I won't venture there).</p></li>\n</ol>\n\n<p>Let me elaborate on the latter, since the former is (probably) already well-known to most haskell programmers here (other than being significantly less relevant):</p>\n\n<p>First we'll need to properly define <span class=\"math-container\">$\\boxtimes$</span>: if we fix functors <span class=\"math-container\">$\\mathcal{F}, \\mathcal{G} : \\mathcal{C} \\to \\mathcal{C}$</span> we can define <span class=\"math-container\">$(\\mathcal{F} \\boxtimes \\mathcal{G}) (a) := \\mathcal{F}(a) \\otimes \\mathcal{G}(a)$</span> (the definition for morphisms is analogous). This, together with <span class=\"math-container\">$\\Delta_I$</span> as a unit, defines a monoidal category: all axioms (and further properties, such as symmetry and strictness) follow trivially from the monoidal structure on <span class=\"math-container\">$\\mathcal{C}$</span>. </p>\n\n<p>Now, monoids in this monoidal category are what haskell programmers call \"alternative\" functors, but that's not all: they are so in a functorial way, meaning that if <span class=\"math-container\">$f : a \\to b : \\mathcal{C}$</span> and <span class=\"math-container\">$\\mathcal{M}$</span> is a monoid in <span class=\"math-container\">$([\\mathcal{C}, \\mathcal{C}], \\boxtimes, \\Delta_I)$</span> then <span class=\"math-container\">$\\mathcal{M}(f)$</span> is a morphism of monoids. To see this, note how <a href=\"https://tikzcd.yichuanshen.de/#N4Igdg9gJgpgziAXAbVABwnAlgFyxMJZABgBpiBdUkANwEMAbAVxiRAB12BbOnACwDGjYAFkAvgAJOAIwgAPPF3hT2AERgMcdAPoBJEGNLpMufIRQBGclVqMWbTj35CGosQaMgM2PASIAma2p6ZlZEDjUNLT0VWQUsJTgVJ0FhcQ9jHzMiMgsbEPtwx15U13FY+UVlYuc090NM0z9LUjzguzCIlJc3DK8TX3NkQLbbUIduEp7ymUqE6snasvqbGCgAc3giUAAzACcILiQyEBwIJCsxwoi9vnPqBjppDQAFAezwvax1vhw+-cOSECp3OiEuBU6nEeXGkUDoIAeT1e72aIAYMB2fwaIABR0QwLOSAArO1xkV2DA0NgGAQKvFEskStIdsALGJtDVSr1Ec8GG8sqivj8sZ5ccdqITEABmHnIgXmNEYv6k66c5ms9mc6aSWb0haU6l+bFi6US0EAFllfJRCvRmIRV0h3CY-wOeJJIKQlrRSOt8rYQt+DohEy4LuNbouZq9KqdKXVbI5iy56TEFDEQA\" rel=\"nofollow noreferrer\">this diagram</a> needs to commute in a pointwise fashion, and the same applies to <a href=\"https://tikzcd.yichuanshen.de/#N4Igdg9gJgpgziAXAbVABwnAlgFyxMJZABgBpiBdUkANwEMAbAVxiRAB12BbOnACwDGjYAFkAvgAJOAIwgAPPF3hTuvQcPErZCrErgqe-IQ1FiQY0uky58hFGQCMVWoxZtOh9Sc0z5i5R5qxqbmliAY2HgERABM5M70zKyIHKpGGpK+OnoGQRmhVpG2saRO1IluKYHp3mZizjBQAObwRKAAZgBOEFxIZCA4EEgO5a7JqVxMWn66AWl80u3ADmIA+tVeIRYd3b2IcQNDiADM1Ax00jAMAArWUXYgnVhNfDggo0nu3EwFIF09w2ogyQpxcnyq31+-z2-WB+zOFyutyK0RSTxebw+lQmakWyzWG2CPnY2n8+g8P3qYiAA\" rel=\"nofollow noreferrer\">this other diagram</a> too. We're in the strict case, but nothing significant changes in the general one (except that you need to keep track of the coherence laws).</p>\n\n<p>A more abstract (but equivalent) formulation is that each object <span class=\"math-container\">$\\mathcal{M} : \\mathbf{Mon}([\\mathcal{C}, \\mathcal{C}])$</span> naturally defines a functor <span class=\"math-container\">$\\hat{\\mathcal{M}} : \\mathcal{C} \\to \\mathbf{Mon}(\\mathcal{C})$</span>, and this point of view leads to the observation that this actually defines a functor <span class=\"math-container\">$(\\hat{-}) : \\mathbf{Mon}([\\mathcal{C}, \\mathcal{C}]) \\to [\\mathcal{C}, \\mathbf{Mon}(\\mathcal{C})]$</span>. </p>\n\n<p>What this amounts to is just that a \"monoidal\" natural transformation <span class=\"math-container\">$\\nu : \\mathcal{M} \\to \\mathcal{N} : \\mathbf{Mon}([\\mathcal{C}, \\mathcal{C}])$</span> between monoid object in this \"pointwise\" monoidal category of (endo)functors also defines a natural transformation <span class=\"math-container\">$\\hat{\\nu} : \\hat{\\mathcal{M}} \\to \\hat{\\mathcal{N}} : [\\mathcal{C}, \\mathbf{Mon}(\\mathcal{C})]$</span>: the argument is similar but you need to look at the diagrams for <a href=\"https://ncatlab.org/nlab/show/monoid+in+a+monoidal+category#morphism_of_monoids\" rel=\"nofollow noreferrer\">morphisms of monoids</a>.</p>\n\n<p>Things are not over yet: it seems that <span class=\"math-container\">$(\\hat{-})$</span> is actually an equivalence of categories! You can convince yourself of this if you squint hard enough, and the argument is roughly \"brute force\":</p>\n\n<p>Let's construct a strict inverse to <span class=\"math-container\">$(\\hat{-})$</span>, call it <span class=\"math-container\">$(\\tilde{-})$</span>: take <span class=\"math-container\">$\\mathcal{F} : \\mathcal{C} \\to \\mathbf{Mon}(\\mathcal{C})$</span>, compose it with the forgetful functor <span class=\"math-container\">$\\mathbf{U} : \\mathbf{Mon}(\\mathcal{C}) \\to \\mathcal{C}$</span>. We'll now use the \"forgotten structure\" on <span class=\"math-container\">$\\mathcal{F}$</span> to endow <span class=\"math-container\">$\\mathbf{U} \\circ \\mathcal{F}$</span> with the structure of a monoid in <span class=\"math-container\">$[\\mathcal{C}, \\mathcal{C}]$</span>: to do this, notice how the monoid structure on <span class=\"math-container\">$\\mathcal{F}(a)$</span> (for <span class=\"math-container\">$a : \\mathcal{C}$</span>) needs to be such that each morphism <span class=\"math-container\">$\\mathcal{F}(f) : \\mathcal{F}(a) \\to \\mathcal{F}(b)$</span> is a morphism of monoids: this amounts to saying that the correspondences <span class=\"math-container\">$a \\mapsto \\mathbf{m}^{\\mathcal{F}(a)}$</span> and <span class=\"math-container\">$a \\mapsto \\mathbf{e}^{\\mathcal{F}(a)}$</span> are natural in <span class=\"math-container\">$a$</span> and hence natural transformations (its now trivial to check that these endow <span class=\"math-container\">$\\mathbf{U} \\circ \\mathcal{F}$</span> with the structure of a monoid in <span class=\"math-container\">$[\\mathcal{C}, \\mathcal{C}]$</span>).</p>\n\n<p>We now need to check that <span class=\"math-container\">$(\\hat{-}) \\circ (\\tilde{-}) = \\mathbf{1}_{[\\mathcal{C}, \\mathbf{Mon}(\\mathcal{C})]}$</span> and <span class=\"math-container\">$(\\tilde{-} \\circ \\hat{-}) = \\mathbf{1}_{\\mathbf{Mon}([\\mathcal{C}, \\mathcal{C}])}$</span>: they should both be long and trivial computations, and I'll (boldly) omit them.</p>\n\n<p>Edit: this can probably be \"merged\" to <a href=\"https://mathoverflow.net/a/259163/111265\">the other answer by Andrej Bauer</a> to give a more complete answer</p>\n" } ]
Question: What is the mathematical name for Haskell&#39;s Alternative Functor Answer 1: <p>This is not a complete answer, just a translation of Haskell to math so that the resident category theorists can tell what the question is.</p> <p>As is usual in such cases, the terminology and the concepts in Haskell are slightly modified concepts from mathematics. It is best in the first iteration to look just at the broad similarities and not worry so much about the details.</p> <p>Recall from <a href="http://www.staff.city.ac.uk/~ross/papers/Applicative.pdf" rel="noreferrer">Section 7 of the paper</a> by Connor McBride and Ross Paterson that Haskell's <code>Applicative</code> is a lax monoidal functor (kind of as the Haskell definition uses internal Hom-sets where one would expect the external ones).</p> <p>The definition of <code>Alternative</code> translates into: a functor $F : \mathcal{C} \to \mathcal{C}$, where $C$ is at least cartesian, together with natural transformations $e : 1 \to F$ and $m : F({-}) \times F({-}) \to F$, or concretely, for every object $A \in \mathcal{C}$ we have arrows $$e_A : 1 \to F(A)$$ and $$m_A : F(A) \times F(A) \to F(A)$$ such that $(F(A), e_A, m_A)$ is a monoid for all $A$, naturally in $A$. What would you call such a thing?</p> <p>The next question to ask is whether there is anything extra about having a lax monoidal functor which also has the above structure of monoids? I don't know off the top of my head. If there is another view of the same situation, it will surely be useful for various Haskell hacking tricks. Haskell people are very good at using category-theoretic algebra for all sorts of cool purposes.</p> <p>Let me also explain on the difference between parametricity and naturality. When we define a "functor" $F$ in Haskell, that is not really a functor. It is acertain mapping from types to types which acts on internal hom-sets as a functor would. In addition, this mapping has a strong property known as <em><a href="https://en.wikipedia.org/wiki/Parametricity" rel="noreferrer">parametricity</a></em> which says that $F$ behaves "in the same way" on all types. Parametricity is quite similar to naturality, and there is a way to view it semantically as such, but it is not precisely naturality. This is not a detail into which we should worry about here, though.</p> Answer 2: <p>A slightly different set of primitive operations that we can look at is the following:</p> <pre><code>class Functor f =&gt; Alternative' f where unit :: () -&gt; f Void zip :: (f a, f b) -&gt; f (Either a b) </code></pre> <p>This represents a lax monoidal functor from the monoidal category of <code>-&gt;</code> under the tensor <code>Either</code> to the monoidal category of <code>-&gt;</code> under the tensor <code>(,)</code>. The laws are the usual laws for the coherence maps of lax monoidal functors, including those that demand conformance with the functor (i.e. you can crib them with minor adjustments from the <code>Applicative</code> typeclass).</p> <p>By contrast, the operations and laws of the <code>Alternative</code> typeclass of today seem to amount to <code>type Alternative f = Monoid (forall a. f a)</code>. We can witness a pair of maps that convert between the primitive operations of <code>Alternative'</code> and <code>Alternative</code>:</p> <pre><code>instance Alternative f =&gt; Alternative' f where unit _ = empty zip (fa, fb) = (Left &lt;<span class="math-container">$&gt; fa) &lt;|&gt; (Right &lt;$</span>&gt; fb) instance Alternative' f =&gt; Alternative f where empty = absurd &lt;<span class="math-container">$&gt; unit () fa1 &lt;|&gt; fa2 = either id id &lt;$</span>&gt; zip (fa1, fa2) </code></pre> <p>However, it seems likely that this is not truly an isomorphism, due to the difference in constraints and laws between the two typeclasses.</p> <p>On one hand, <code>Alternative</code> is more permissive than <code>Alternative'</code>, since it does not demand any form of compatibility between the monoid and the functor, whereas the commutative diagrams of a lax monoidal functor do.</p> <p>On, the other hand, <code>Alternative</code> demands <code>Applicative</code>, a constraint that <code>Alternative'</code> does not demand. When viewed through the lens of lax monoidal functors, <code>Applicative f</code> simply means that <code>f</code> coheres the <code>(,)</code> monoidal structure to <code>(,)</code>, and <code>Alternative f</code> means that it coheres the <code>Either</code> monoidal structure to <code>(,)</code>. These are orthogonal claims, so there is no need for <code>Alternative' f</code> to demand a <code>Applicative f</code> constraint.</p> <p>That said, I conjecture that most types that support <code>Alternative</code> instances also support a lawful instance of <code>Alternative'</code>, and so you can imagine them to be "applicatives for different tensor".</p> Answer 3: <p>This might not be the right approach, but here's my take on the question:</p> <p>If <span class="math-container">$(\mathcal{C}, \otimes, I, \dots)$</span> is a (strict, for the sake of readability) monoidal category, we can endow <span class="math-container">$[\mathcal{C}, \mathcal{C}]$</span> with two different monoidal structures:</p> <ol> <li><p>The "usual" monoidal structure, where <span class="math-container">$\circ$</span> is the product and <span class="math-container">$\mathbf{1}_\mathcal{C}$</span> is the unit. This is the monoidal category whose monoids are called monads. Notice how this doesn't actually use the fact that <span class="math-container">$\mathcal{C}$</span> is endowed with a monoidal structure.</p></li> <li><p>The "other" (I sometimes refer to it as "pointwise") monoidal structure has "pointwise <span class="math-container">$\otimes$</span>" (let's call it <span class="math-container">$\boxtimes$</span>) as product, and <span class="math-container">$\Delta_I$</span> (the constant functor on the unit object of <span class="math-container">$\mathcal{C}$</span>) as monoidal units. In this category, monoids are related to what Haskell programmers call "alternative" functors, but without the part about being "applicative" and with one additional property. This can be further generalized to the case of functors <span class="math-container">$\mathcal{B} \to \mathcal{C}$</span> where <span class="math-container">$\mathcal{B}$</span> is <em>not</em> required to be monoidal, and everything that follows should work just fine (but I won't venture there).</p></li> </ol> <p>Let me elaborate on the latter, since the former is (probably) already well-known to most haskell programmers here (other than being significantly less relevant):</p> <p>First we'll need to properly define <span class="math-container">$\boxtimes$</span>: if we fix functors <span class="math-container">$\mathcal{F}, \mathcal{G} : \mathcal{C} \to \mathcal{C}$</span> we can define <span class="math-container">$(\mathcal{F} \boxtimes \mathcal{G}) (a) := \mathcal{F}(a) \otimes \mathcal{G}(a)$</span> (the definition for morphisms is analogous). This, together with <span class="math-container">$\Delta_I$</span> as a unit, defines a monoidal category: all axioms (and further properties, such as symmetry and strictness) follow trivially from the monoidal structure on <span class="math-container">$\mathcal{C}$</span>. </p> <p>Now, monoids in this monoidal category are what haskell programmers call "alternative" functors, but that's not all: they are so in a functorial way, meaning that if <span class="math-container">$f : a \to b : \mathcal{C}$</span> and <span class="math-container">$\mathcal{M}$</span> is a monoid in <span class="math-container">$([\mathcal{C}, \mathcal{C}], \boxtimes, \Delta_I)$</span> then <span class="math-container">$\mathcal{M}(f)$</span> is a morphism of monoids. To see this, note how <a href="https://tikzcd.yichuanshen.de/#N4Igdg9gJgpgziAXAbVABwnAlgFyxMJZABgBpiBdUkANwEMAbAVxiRAB12BbOnACwDGjYAFkAvgAJOAIwgAPPF3hT2AERgMcdAPoBJEGNLpMufIRQBGclVqMWbTj35CGosQaMgM2PASIAma2p6ZlZEDjUNLT0VWQUsJTgVJ0FhcQ9jHzMiMgsbEPtwx15U13FY+UVlYuc090NM0z9LUjzguzCIlJc3DK8TX3NkQLbbUIduEp7ymUqE6snasvqbGCgAc3giUAAzACcILiQyEBwIJCsxwoi9vnPqBjppDQAFAezwvax1vhw+-cOSECp3OiEuBU6nEeXGkUDoIAeT1e72aIAYMB2fwaIABR0QwLOSAArO1xkV2DA0NgGAQKvFEskStIdsALGJtDVSr1Ec8GG8sqivj8sZ5ccdqITEABmHnIgXmNEYv6k66c5ms9mc6aSWb0haU6l+bFi6US0EAFllfJRCvRmIRV0h3CY-wOeJJIKQlrRSOt8rYQt+DohEy4LuNbouZq9KqdKXVbI5iy56TEFDEQA" rel="nofollow noreferrer">this diagram</a> needs to commute in a pointwise fashion, and the same applies to <a href="https://tikzcd.yichuanshen.de/#N4Igdg9gJgpgziAXAbVABwnAlgFyxMJZABgBpiBdUkANwEMAbAVxiRAB12BbOnACwDGjYAFkAvgAJOAIwgAPPF3hTuvQcPErZCrErgqe-IQ1FiQY0uky58hFGQCMVWoxZtOh9Sc0z5i5R5qxqbmliAY2HgERABM5M70zKyIHKpGGpK+OnoGQRmhVpG2saRO1IluKYHp3mZizjBQAObwRKAAZgBOEFxIZCA4EEgO5a7JqVxMWn66AWl80u3ADmIA+tVeIRYd3b2IcQNDiADM1Ax00jAMAArWUXYgnVhNfDggo0nu3EwFIF09w2ogyQpxcnyq31+-z2-WB+zOFyutyK0RSTxebw+lQmakWyzWG2CPnY2n8+g8P3qYiAA" rel="nofollow noreferrer">this other diagram</a> too. We're in the strict case, but nothing significant changes in the general one (except that you need to keep track of the coherence laws).</p> <p>A more abstract (but equivalent) formulation is that each object <span class="math-container">$\mathcal{M} : \mathbf{Mon}([\mathcal{C}, \mathcal{C}])$</span> naturally defines a functor <span class="math-container">$\hat{\mathcal{M}} : \mathcal{C} \to \mathbf{Mon}(\mathcal{C})$</span>, and this point of view leads to the observation that this actually defines a functor <span class="math-container">$(\hat{-}) : \mathbf{Mon}([\mathcal{C}, \mathcal{C}]) \to [\mathcal{C}, \mathbf{Mon}(\mathcal{C})]$</span>. </p> <p>What this amounts to is just that a "monoidal" natural transformation <span class="math-container">$\nu : \mathcal{M} \to \mathcal{N} : \mathbf{Mon}([\mathcal{C}, \mathcal{C}])$</span> between monoid object in this "pointwise" monoidal category of (endo)functors also defines a natural transformation <span class="math-container">$\hat{\nu} : \hat{\mathcal{M}} \to \hat{\mathcal{N}} : [\mathcal{C}, \mathbf{Mon}(\mathcal{C})]$</span>: the argument is similar but you need to look at the diagrams for <a href="https://ncatlab.org/nlab/show/monoid+in+a+monoidal+category#morphism_of_monoids" rel="nofollow noreferrer">morphisms of monoids</a>.</p> <p>Things are not over yet: it seems that <span class="math-container">$(\hat{-})$</span> is actually an equivalence of categories! You can convince yourself of this if you squint hard enough, and the argument is roughly "brute force":</p> <p>Let's construct a strict inverse to <span class="math-container">$(\hat{-})$</span>, call it <span class="math-container">$(\tilde{-})$</span>: take <span class="math-container">$\mathcal{F} : \mathcal{C} \to \mathbf{Mon}(\mathcal{C})$</span>, compose it with the forgetful functor <span class="math-container">$\mathbf{U} : \mathbf{Mon}(\mathcal{C}) \to \mathcal{C}$</span>. We'll now use the "forgotten structure" on <span class="math-container">$\mathcal{F}$</span> to endow <span class="math-container">$\mathbf{U} \circ \mathcal{F}$</span> with the structure of a monoid in <span class="math-container">$[\mathcal{C}, \mathcal{C}]$</span>: to do this, notice how the monoid structure on <span class="math-container">$\mathcal{F}(a)$</span> (for <span class="math-container">$a : \mathcal{C}$</span>) needs to be such that each morphism <span class="math-container">$\mathcal{F}(f) : \mathcal{F}(a) \to \mathcal{F}(b)$</span> is a morphism of monoids: this amounts to saying that the correspondences <span class="math-container">$a \mapsto \mathbf{m}^{\mathcal{F}(a)}$</span> and <span class="math-container">$a \mapsto \mathbf{e}^{\mathcal{F}(a)}$</span> are natural in <span class="math-container">$a$</span> and hence natural transformations (its now trivial to check that these endow <span class="math-container">$\mathbf{U} \circ \mathcal{F}$</span> with the structure of a monoid in <span class="math-container">$[\mathcal{C}, \mathcal{C}]$</span>).</p> <p>We now need to check that <span class="math-container">$(\hat{-}) \circ (\tilde{-}) = \mathbf{1}_{[\mathcal{C}, \mathbf{Mon}(\mathcal{C})]}$</span> and <span class="math-container">$(\tilde{-} \circ \hat{-}) = \mathbf{1}_{\mathbf{Mon}([\mathcal{C}, \mathcal{C}])}$</span>: they should both be long and trivial computations, and I'll (boldly) omit them.</p> <p>Edit: this can probably be "merged" to <a href="https://mathoverflow.net/a/259163/111265">the other answer by Andrej Bauer</a> to give a more complete answer</p> Answer 4: <p>This is not a complete answer, just a translation of Haskell to math so that the resident category theorists can tell what the question is.</p> <p>As is usual in such cases, the terminology and the concepts in Haskell are slightly modified concepts from mathematics. It is best in the first iteration to look just at the broad similarities and not worry so much about the details.</p> <p>Recall from <a href="http://www.staff.city.ac.uk/~ross/papers/Applicative.pdf" rel="noreferrer">Section 7 of the paper</a> by Connor McBride and Ross Paterson that Haskell's <code>Applicative</code> is a lax monoidal functor (kind of as the Haskell definition uses internal Hom-sets where one would expect the external ones).</p> <p>The definition of <code>Alternative</code> translates into: a functor $F : \mathcal{C} \to \mathcal{C}$, where $C$ is at least cartesian, together with natural transformations $e : 1 \to F$ and $m : F({-}) \times F({-}) \to F$, or concretely, for every object $A \in \mathcal{C}$ we have arrows $$e_A : 1 \to F(A)$$ and $$m_A : F(A) \times F(A) \to F(A)$$ such that $(F(A), e_A, m_A)$ is a monoid for all $A$, naturally in $A$. What would you call such a thing?</p> <p>The next question to ask is whether there is anything extra about having a lax monoidal functor which also has the above structure of monoids? I don't know off the top of my head. If there is another view of the same situation, it will surely be useful for various Haskell hacking tricks. Haskell people are very good at using category-theoretic algebra for all sorts of cool purposes.</p> <p>Let me also explain on the difference between parametricity and naturality. When we define a "functor" $F$ in Haskell, that is not really a functor. It is acertain mapping from types to types which acts on internal hom-sets as a functor would. In addition, this mapping has a strong property known as <em><a href="https://en.wikipedia.org/wiki/Parametricity" rel="noreferrer">parametricity</a></em> which says that $F$ behaves "in the same way" on all types. Parametricity is quite similar to naturality, and there is a way to view it semantically as such, but it is not precisely naturality. This is not a detail into which we should worry about here, though.</p> Answer 5: <p>A slightly different set of primitive operations that we can look at is the following:</p> <pre><code>class Functor f =&gt; Alternative' f where unit :: () -&gt; f Void zip :: (f a, f b) -&gt; f (Either a b) </code></pre> <p>This represents a lax monoidal functor from the monoidal category of <code>-&gt;</code> under the tensor <code>Either</code> to the monoidal category of <code>-&gt;</code> under the tensor <code>(,)</code>. The laws are the usual laws for the coherence maps of lax monoidal functors, including those that demand conformance with the functor (i.e. you can crib them with minor adjustments from the <code>Applicative</code> typeclass).</p> <p>By contrast, the operations and laws of the <code>Alternative</code> typeclass of today seem to amount to <code>type Alternative f = Monoid (forall a. f a)</code>. We can witness a pair of maps that convert between the primitive operations of <code>Alternative'</code> and <code>Alternative</code>:</p> <pre><code>instance Alternative f =&gt; Alternative' f where unit _ = empty zip (fa, fb) = (Left &lt;<span class="math-container">$&gt; fa) &lt;|&gt; (Right &lt;$</span>&gt; fb) instance Alternative' f =&gt; Alternative f where empty = absurd &lt;<span class="math-container">$&gt; unit () fa1 &lt;|&gt; fa2 = either id id &lt;$</span>&gt; zip (fa1, fa2) </code></pre> <p>However, it seems likely that this is not truly an isomorphism, due to the difference in constraints and laws between the two typeclasses.</p> <p>On one hand, <code>Alternative</code> is more permissive than <code>Alternative'</code>, since it does not demand any form of compatibility between the monoid and the functor, whereas the commutative diagrams of a lax monoidal functor do.</p> <p>On, the other hand, <code>Alternative</code> demands <code>Applicative</code>, a constraint that <code>Alternative'</code> does not demand. When viewed through the lens of lax monoidal functors, <code>Applicative f</code> simply means that <code>f</code> coheres the <code>(,)</code> monoidal structure to <code>(,)</code>, and <code>Alternative f</code> means that it coheres the <code>Either</code> monoidal structure to <code>(,)</code>. These are orthogonal claims, so there is no need for <code>Alternative' f</code> to demand a <code>Applicative f</code> constraint.</p> <p>That said, I conjecture that most types that support <code>Alternative</code> instances also support a lawful instance of <code>Alternative'</code>, and so you can imagine them to be "applicatives for different tensor".</p> Answer 6: <p>This might not be the right approach, but here's my take on the question:</p> <p>If <span class="math-container">$(\mathcal{C}, \otimes, I, \dots)$</span> is a (strict, for the sake of readability) monoidal category, we can endow <span class="math-container">$[\mathcal{C}, \mathcal{C}]$</span> with two different monoidal structures:</p> <ol> <li><p>The "usual" monoidal structure, where <span class="math-container">$\circ$</span> is the product and <span class="math-container">$\mathbf{1}_\mathcal{C}$</span> is the unit. This is the monoidal category whose monoids are called monads. Notice how this doesn't actually use the fact that <span class="math-container">$\mathcal{C}$</span> is endowed with a monoidal structure.</p></li> <li><p>The "other" (I sometimes refer to it as "pointwise") monoidal structure has "pointwise <span class="math-container">$\otimes$</span>" (let's call it <span class="math-container">$\boxtimes$</span>) as product, and <span class="math-container">$\Delta_I$</span> (the constant functor on the unit object of <span class="math-container">$\mathcal{C}$</span>) as monoidal units. In this category, monoids are related to what Haskell programmers call "alternative" functors, but without the part about being "applicative" and with one additional property. This can be further generalized to the case of functors <span class="math-container">$\mathcal{B} \to \mathcal{C}$</span> where <span class="math-container">$\mathcal{B}$</span> is <em>not</em> required to be monoidal, and everything that follows should work just fine (but I won't venture there).</p></li> </ol> <p>Let me elaborate on the latter, since the former is (probably) already well-known to most haskell programmers here (other than being significantly less relevant):</p> <p>First we'll need to properly define <span class="math-container">$\boxtimes$</span>: if we fix functors <span class="math-container">$\mathcal{F}, \mathcal{G} : \mathcal{C} \to \mathcal{C}$</span> we can define <span class="math-container">$(\mathcal{F} \boxtimes \mathcal{G}) (a) := \mathcal{F}(a) \otimes \mathcal{G}(a)$</span> (the definition for morphisms is analogous). This, together with <span class="math-container">$\Delta_I$</span> as a unit, defines a monoidal category: all axioms (and further properties, such as symmetry and strictness) follow trivially from the monoidal structure on <span class="math-container">$\mathcal{C}$</span>. </p> <p>Now, monoids in this monoidal category are what haskell programmers call "alternative" functors, but that's not all: they are so in a functorial way, meaning that if <span class="math-container">$f : a \to b : \mathcal{C}$</span> and <span class="math-container">$\mathcal{M}$</span> is a monoid in <span class="math-container">$([\mathcal{C}, \mathcal{C}], \boxtimes, \Delta_I)$</span> then <span class="math-container">$\mathcal{M}(f)$</span> is a morphism of monoids. To see this, note how <a href="https://tikzcd.yichuanshen.de/#N4Igdg9gJgpgziAXAbVABwnAlgFyxMJZABgBpiBdUkANwEMAbAVxiRAB12BbOnACwDGjYAFkAvgAJOAIwgAPPF3hT2AERgMcdAPoBJEGNLpMufIRQBGclVqMWbTj35CGosQaMgM2PASIAma2p6ZlZEDjUNLT0VWQUsJTgVJ0FhcQ9jHzMiMgsbEPtwx15U13FY+UVlYuc090NM0z9LUjzguzCIlJc3DK8TX3NkQLbbUIduEp7ymUqE6snasvqbGCgAc3giUAAzACcILiQyEBwIJCsxwoi9vnPqBjppDQAFAezwvax1vhw+-cOSECp3OiEuBU6nEeXGkUDoIAeT1e72aIAYMB2fwaIABR0QwLOSAArO1xkV2DA0NgGAQKvFEskStIdsALGJtDVSr1Ec8GG8sqivj8sZ5ccdqITEABmHnIgXmNEYv6k66c5ms9mc6aSWb0haU6l+bFi6US0EAFllfJRCvRmIRV0h3CY-wOeJJIKQlrRSOt8rYQt+DohEy4LuNbouZq9KqdKXVbI5iy56TEFDEQA" rel="nofollow noreferrer">this diagram</a> needs to commute in a pointwise fashion, and the same applies to <a href="https://tikzcd.yichuanshen.de/#N4Igdg9gJgpgziAXAbVABwnAlgFyxMJZABgBpiBdUkANwEMAbAVxiRAB12BbOnACwDGjYAFkAvgAJOAIwgAPPF3hTuvQcPErZCrErgqe-IQ1FiQY0uky58hFGQCMVWoxZtOh9Sc0z5i5R5qxqbmliAY2HgERABM5M70zKyIHKpGGpK+OnoGQRmhVpG2saRO1IluKYHp3mZizjBQAObwRKAAZgBOEFxIZCA4EEgO5a7JqVxMWn66AWl80u3ADmIA+tVeIRYd3b2IcQNDiADM1Ax00jAMAArWUXYgnVhNfDggo0nu3EwFIF09w2ogyQpxcnyq31+-z2-WB+zOFyutyK0RSTxebw+lQmakWyzWG2CPnY2n8+g8P3qYiAA" rel="nofollow noreferrer">this other diagram</a> too. We're in the strict case, but nothing significant changes in the general one (except that you need to keep track of the coherence laws).</p> <p>A more abstract (but equivalent) formulation is that each object <span class="math-container">$\mathcal{M} : \mathbf{Mon}([\mathcal{C}, \mathcal{C}])$</span> naturally defines a functor <span class="math-container">$\hat{\mathcal{M}} : \mathcal{C} \to \mathbf{Mon}(\mathcal{C})$</span>, and this point of view leads to the observation that this actually defines a functor <span class="math-container">$(\hat{-}) : \mathbf{Mon}([\mathcal{C}, \mathcal{C}]) \to [\mathcal{C}, \mathbf{Mon}(\mathcal{C})]$</span>. </p> <p>What this amounts to is just that a "monoidal" natural transformation <span class="math-container">$\nu : \mathcal{M} \to \mathcal{N} : \mathbf{Mon}([\mathcal{C}, \mathcal{C}])$</span> between monoid object in this "pointwise" monoidal category of (endo)functors also defines a natural transformation <span class="math-container">$\hat{\nu} : \hat{\mathcal{M}} \to \hat{\mathcal{N}} : [\mathcal{C}, \mathbf{Mon}(\mathcal{C})]$</span>: the argument is similar but you need to look at the diagrams for <a href="https://ncatlab.org/nlab/show/monoid+in+a+monoidal+category#morphism_of_monoids" rel="nofollow noreferrer">morphisms of monoids</a>.</p> <p>Things are not over yet: it seems that <span class="math-container">$(\hat{-})$</span> is actually an equivalence of categories! You can convince yourself of this if you squint hard enough, and the argument is roughly "brute force":</p> <p>Let's construct a strict inverse to <span class="math-container">$(\hat{-})$</span>, call it <span class="math-container">$(\tilde{-})$</span>: take <span class="math-container">$\mathcal{F} : \mathcal{C} \to \mathbf{Mon}(\mathcal{C})$</span>, compose it with the forgetful functor <span class="math-container">$\mathbf{U} : \mathbf{Mon}(\mathcal{C}) \to \mathcal{C}$</span>. We'll now use the "forgotten structure" on <span class="math-container">$\mathcal{F}$</span> to endow <span class="math-container">$\mathbf{U} \circ \mathcal{F}$</span> with the structure of a monoid in <span class="math-container">$[\mathcal{C}, \mathcal{C}]$</span>: to do this, notice how the monoid structure on <span class="math-container">$\mathcal{F}(a)$</span> (for <span class="math-container">$a : \mathcal{C}$</span>) needs to be such that each morphism <span class="math-container">$\mathcal{F}(f) : \mathcal{F}(a) \to \mathcal{F}(b)$</span> is a morphism of monoids: this amounts to saying that the correspondences <span class="math-container">$a \mapsto \mathbf{m}^{\mathcal{F}(a)}$</span> and <span class="math-container">$a \mapsto \mathbf{e}^{\mathcal{F}(a)}$</span> are natural in <span class="math-container">$a$</span> and hence natural transformations (its now trivial to check that these endow <span class="math-container">$\mathbf{U} \circ \mathcal{F}$</span> with the structure of a monoid in <span class="math-container">$[\mathcal{C}, \mathcal{C}]$</span>).</p> <p>We now need to check that <span class="math-container">$(\hat{-}) \circ (\tilde{-}) = \mathbf{1}_{[\mathcal{C}, \mathbf{Mon}(\mathcal{C})]}$</span> and <span class="math-container">$(\tilde{-} \circ \hat{-}) = \mathbf{1}_{\mathbf{Mon}([\mathcal{C}, \mathcal{C}])}$</span>: they should both be long and trivial computations, and I'll (boldly) omit them.</p> <p>Edit: this can probably be "merged" to <a href="https://mathoverflow.net/a/259163/111265">the other answer by Andrej Bauer</a> to give a more complete answer</p>
Combinatorial interpretation of series reversion coefficients
259,374
8
https://mathoverflow.net/questions/259374/combinatorial-interpretation-of-series-reversion-coefficients
<p>In a recent <a href="https://arxiv.org/abs/1610.05803" rel="noreferrer">paper</a> studying some generalizations of Stirling numbers, my coauthors and I needed the following result:</p> <p>If $f(x)=\sum_{n \geq 1} a_n x^n/n!$ is a power series with $a_1 \neq 0$, and $g(x) = \sum_{n \geq 1} b_n x^n/n!$ is its series reversion (so $f(g(x))=g(f(x))=x$), then $$ b_n = \sum_{T} w(T) $$<br> where the sum is over phylogenetic trees on $\{1, \ldots, n\}$. Here a <em>phylogenetic tree</em> on $\{1,\ldots, n\}$ is a rooted tree with $n$ leaves, no vertex with exactly one child, and leaves labelled from the set $\{1, \ldots, n\}$ (but no other vertex receiving a label); see page 3 of the paper linked above for a picture. The weight $w(T)$ is computed as follows: if $T$ has $n$ leaves and $m$ non-leaves then $$ w(T) = (−1)^m a_1^{-(m+n)} \prod \{a_{d(v)}: v~\mbox{a non-leaf}\} $$ where $d(v)$ denotes the number of children of $v$ (and when $n=1$ the isolated vertex is considered a leaf).</p> <p>The proof is short, and this feels like something that should have been written down prior to now, but we could find no reference.</p> <p><strong>Question</strong>: Has the above result appeared in the literature?</p> <p>(There is a <a href="https://pdfs.semanticscholar.org/b079/4b05709432e882698c47fbeac625d291743d.pdf" rel="noreferrer">paper</a> of Chen from 1990 with a similar result but with a distinctly different and not obviously equivalent formulation. Chen shows that $b_n$ is a weighted sum over <em>Schroeder trees</em>: trees on $n$ labelled <em>vertices</em> (as opposed to leaves), with no restriction on numbers of children, in which each non-leaf is endowed with an ordered partition of its children. The count of Schroeder trees by number of vertices is <a href="http://oeis.org/A053492" rel="noreferrer">A053492</a>. The count of phylogenetic trees by number of leaves is <a href="http://oeis.org/A000311" rel="noreferrer">A000311</a>.) </p>
[ { "answer_id": 259463, "score": 6, "text": "<p>Yes, this is known. As Tom Copeland mentioned, this is in <a href=\"http://people.brandeis.edu/~gessel/homepage/students/drakethesis.pdf\" rel=\"nofollow noreferrer\">Drake's thesis</a>, Example 1.4.7.</p>\n\n<p>Drake gives a broad generalization of your result. His main theorem (Theorem 1.3.3) is roughly as follows. Suppose you have an alphabet $A$ of certain trees, and a set $L$ of 'allowed links' between them, ways of connecting the root of one to a leaf of another. Let $S(L,n)$ be the set of all trees $T$ with $n$ leaves labeled bijectively $1, \\ldots, n$ with no other nodes labeled with the property that $T$ made up of the letters $A$ with <em>only</em> links between them which are in $L$. Say each tree $T \\in A$ has a weight $w(T)$, and a tree made up of letters $T_1, \\ldots, T_n \\in A$ with certain links between them has weight $w(T_1) \\cdots w(T_n)$. Let $$F(x) = \\sum_{n=1}^\\infty a_n \\frac{x^n}{n!}$$ where $a_n = \\sum_{T \\in S(L,n)} w(T)$. Then the compositional inverse is $$F^{-1}(x) = \\sum_{n=1}^\\infty b_n \\frac{x^n}{n!}$$ where $$b_n = \\sum_{T \\in S(\\bar{L},n)} (-1)^{m(T)} w(T).$$\nHere $m(T)$ is the number of letters in $A$ that $T$ is composed of, and $\\bar{L}$ is the complement of $L$ in the set of all possible links between elements of $A$.</p>\n\n<p>This was an extension of the work of Susan Parker, who had a similar theorem using ordinary generating functions. Parker's result can, in turn, be seen as a generalization of a result of Carlitz, Scoville and Vaughan. Ira Gessel's <a href=\"http://people.brandeis.edu/~gessel/homepage/slides/csv.pdf\" rel=\"nofollow noreferrer\">slides</a> may be helpful.</p>\n\n<p>To get your theorem, let $A$ be the set of \"simple\" trees, the trees $T_n$, $n &gt; 1$, so that $T_n$ is a tree with the root having the $n$ leaves $1,2, \\ldots, n$ as children. Let $w(T_n) = t_n$ (an indeterminate). Let $L$ be the empty set of links. Then $$F(x) = x + t_2\\frac{x^2}{2!} + t_3 \\frac{x^3}{3!} + \\cdots$$ since $S(L,n) = \\{T_n\\}$ for $n&gt;1$, and $S(L,1)$ consists of the tree with one node which is always allowed (and has weight $1$).</p>\n\n<p>Then by Drake's inversion theorem, $$F^{-1}(x) = \\sum_{n=1}^\\infty b_n \\frac{x^n}{n!}$$ where $$b_n = \\sum_{T \\in S(\\bar{L},n)} (-1)^{m(T)} w(T)$$ is the sum of the weights of <em>all</em> phylogenetic trees (since $\\bar{L}$ is the set of all links.)</p>\n\n<p>See also <a href=\"https://digital.lib.washington.edu/researchworks/handle/1773/36757\" rel=\"nofollow noreferrer\">my thesis</a>, which contains a different generalization of your result. The phylogenetic trees (also called total partitions) are mentioned in Section 3.2</p>\n" }, { "answer_id": 259564, "score": 6, "text": "<p>This result can also be found as Theorem 3.3.9 in Ginzburg-Kapranov: \"Koszul duality for operads\" (Duke Math J 1994). Set $r=1$ in their result and rescale so that $a_1=1$. They had discovered it independently, but thank D. Wright for informing them that the result was known before due to J. Towber. They also give a reference to Wright: \"The tree formulas for reversion of power series\" (JPAA 1989), where this result appears as Theorem 3.10.</p>\n\n<p>The context of this result in the setting of operads is that any dg operad $P = \\{P(n)\\}$, say with $P(0)=0$ and $P(1)$ a copy of the base field, has a generating series which encodes the Euler characteristics of the chain complexes $P(n)$. The generating series of $P$ is the compositional inverse of the generating series for the \"bar construction\" on the operad $P$. The bar construction of operads is defined by a sum over trees whose internal vertices are decorated by the chain complexes of the operad $P$; in fact, the trees involved are exactly those that you call phylogenetic trees.</p>\n" }, { "answer_id": 259376, "score": 2, "text": "<p>See these three papers and references therein:</p>\n\n<p><a href=\"https://arxiv.org/abs/math/0208174\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/math/0208174</a></p>\n\n<p><a href=\"https://arxiv.org/abs/math/0208173\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/math/0208173</a></p>\n\n<p><a href=\"http://www.emis.de/journals/SLC/wpapers/s49abdess.html\" rel=\"nofollow noreferrer\">http://www.emis.de/journals/SLC/wpapers/s49abdess.html</a></p>\n\n<p>Going from one type of tree to another is no big deal in the framework of Joyal's theory of combinatorial species. Types of trees correspond to certain functors between finite sets and for each such functors there is an associated exponential generating series. Changing from say planar trees to Cayley trees is a morphism of functors and there is a precise theorem which says what happens then to the generating series (Theorem 8 in the last paper I listed).</p>\n" }, { "answer_id": 259463, "score": 6, "text": "<p>Yes, this is known. As Tom Copeland mentioned, this is in <a href=\"http://people.brandeis.edu/~gessel/homepage/students/drakethesis.pdf\" rel=\"nofollow noreferrer\">Drake's thesis</a>, Example 1.4.7.</p>\n\n<p>Drake gives a broad generalization of your result. His main theorem (Theorem 1.3.3) is roughly as follows. Suppose you have an alphabet $A$ of certain trees, and a set $L$ of 'allowed links' between them, ways of connecting the root of one to a leaf of another. Let $S(L,n)$ be the set of all trees $T$ with $n$ leaves labeled bijectively $1, \\ldots, n$ with no other nodes labeled with the property that $T$ made up of the letters $A$ with <em>only</em> links between them which are in $L$. Say each tree $T \\in A$ has a weight $w(T)$, and a tree made up of letters $T_1, \\ldots, T_n \\in A$ with certain links between them has weight $w(T_1) \\cdots w(T_n)$. Let $$F(x) = \\sum_{n=1}^\\infty a_n \\frac{x^n}{n!}$$ where $a_n = \\sum_{T \\in S(L,n)} w(T)$. Then the compositional inverse is $$F^{-1}(x) = \\sum_{n=1}^\\infty b_n \\frac{x^n}{n!}$$ where $$b_n = \\sum_{T \\in S(\\bar{L},n)} (-1)^{m(T)} w(T).$$\nHere $m(T)$ is the number of letters in $A$ that $T$ is composed of, and $\\bar{L}$ is the complement of $L$ in the set of all possible links between elements of $A$.</p>\n\n<p>This was an extension of the work of Susan Parker, who had a similar theorem using ordinary generating functions. Parker's result can, in turn, be seen as a generalization of a result of Carlitz, Scoville and Vaughan. Ira Gessel's <a href=\"http://people.brandeis.edu/~gessel/homepage/slides/csv.pdf\" rel=\"nofollow noreferrer\">slides</a> may be helpful.</p>\n\n<p>To get your theorem, let $A$ be the set of \"simple\" trees, the trees $T_n$, $n &gt; 1$, so that $T_n$ is a tree with the root having the $n$ leaves $1,2, \\ldots, n$ as children. Let $w(T_n) = t_n$ (an indeterminate). Let $L$ be the empty set of links. Then $$F(x) = x + t_2\\frac{x^2}{2!} + t_3 \\frac{x^3}{3!} + \\cdots$$ since $S(L,n) = \\{T_n\\}$ for $n&gt;1$, and $S(L,1)$ consists of the tree with one node which is always allowed (and has weight $1$).</p>\n\n<p>Then by Drake's inversion theorem, $$F^{-1}(x) = \\sum_{n=1}^\\infty b_n \\frac{x^n}{n!}$$ where $$b_n = \\sum_{T \\in S(\\bar{L},n)} (-1)^{m(T)} w(T)$$ is the sum of the weights of <em>all</em> phylogenetic trees (since $\\bar{L}$ is the set of all links.)</p>\n\n<p>See also <a href=\"https://digital.lib.washington.edu/researchworks/handle/1773/36757\" rel=\"nofollow noreferrer\">my thesis</a>, which contains a different generalization of your result. The phylogenetic trees (also called total partitions) are mentioned in Section 3.2</p>\n" }, { "answer_id": 259564, "score": 6, "text": "<p>This result can also be found as Theorem 3.3.9 in Ginzburg-Kapranov: \"Koszul duality for operads\" (Duke Math J 1994). Set $r=1$ in their result and rescale so that $a_1=1$. They had discovered it independently, but thank D. Wright for informing them that the result was known before due to J. Towber. They also give a reference to Wright: \"The tree formulas for reversion of power series\" (JPAA 1989), where this result appears as Theorem 3.10.</p>\n\n<p>The context of this result in the setting of operads is that any dg operad $P = \\{P(n)\\}$, say with $P(0)=0$ and $P(1)$ a copy of the base field, has a generating series which encodes the Euler characteristics of the chain complexes $P(n)$. The generating series of $P$ is the compositional inverse of the generating series for the \"bar construction\" on the operad $P$. The bar construction of operads is defined by a sum over trees whose internal vertices are decorated by the chain complexes of the operad $P$; in fact, the trees involved are exactly those that you call phylogenetic trees.</p>\n" }, { "answer_id": 259376, "score": 2, "text": "<p>See these three papers and references therein:</p>\n\n<p><a href=\"https://arxiv.org/abs/math/0208174\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/math/0208174</a></p>\n\n<p><a href=\"https://arxiv.org/abs/math/0208173\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/math/0208173</a></p>\n\n<p><a href=\"http://www.emis.de/journals/SLC/wpapers/s49abdess.html\" rel=\"nofollow noreferrer\">http://www.emis.de/journals/SLC/wpapers/s49abdess.html</a></p>\n\n<p>Going from one type of tree to another is no big deal in the framework of Joyal's theory of combinatorial species. Types of trees correspond to certain functors between finite sets and for each such functors there is an associated exponential generating series. Changing from say planar trees to Cayley trees is a morphism of functors and there is a precise theorem which says what happens then to the generating series (Theorem 8 in the last paper I listed).</p>\n" } ]
Question: Combinatorial interpretation of series reversion coefficients Answer 1: <p>Yes, this is known. As Tom Copeland mentioned, this is in <a href="http://people.brandeis.edu/~gessel/homepage/students/drakethesis.pdf" rel="nofollow noreferrer">Drake's thesis</a>, Example 1.4.7.</p> <p>Drake gives a broad generalization of your result. His main theorem (Theorem 1.3.3) is roughly as follows. Suppose you have an alphabet $A$ of certain trees, and a set $L$ of 'allowed links' between them, ways of connecting the root of one to a leaf of another. Let $S(L,n)$ be the set of all trees $T$ with $n$ leaves labeled bijectively $1, \ldots, n$ with no other nodes labeled with the property that $T$ made up of the letters $A$ with <em>only</em> links between them which are in $L$. Say each tree $T \in A$ has a weight $w(T)$, and a tree made up of letters $T_1, \ldots, T_n \in A$ with certain links between them has weight $w(T_1) \cdots w(T_n)$. Let $$F(x) = \sum_{n=1}^\infty a_n \frac{x^n}{n!}$$ where $a_n = \sum_{T \in S(L,n)} w(T)$. Then the compositional inverse is $$F^{-1}(x) = \sum_{n=1}^\infty b_n \frac{x^n}{n!}$$ where $$b_n = \sum_{T \in S(\bar{L},n)} (-1)^{m(T)} w(T).$$ Here $m(T)$ is the number of letters in $A$ that $T$ is composed of, and $\bar{L}$ is the complement of $L$ in the set of all possible links between elements of $A$.</p> <p>This was an extension of the work of Susan Parker, who had a similar theorem using ordinary generating functions. Parker's result can, in turn, be seen as a generalization of a result of Carlitz, Scoville and Vaughan. Ira Gessel's <a href="http://people.brandeis.edu/~gessel/homepage/slides/csv.pdf" rel="nofollow noreferrer">slides</a> may be helpful.</p> <p>To get your theorem, let $A$ be the set of "simple" trees, the trees $T_n$, $n &gt; 1$, so that $T_n$ is a tree with the root having the $n$ leaves $1,2, \ldots, n$ as children. Let $w(T_n) = t_n$ (an indeterminate). Let $L$ be the empty set of links. Then $$F(x) = x + t_2\frac{x^2}{2!} + t_3 \frac{x^3}{3!} + \cdots$$ since $S(L,n) = \{T_n\}$ for $n&gt;1$, and $S(L,1)$ consists of the tree with one node which is always allowed (and has weight $1$).</p> <p>Then by Drake's inversion theorem, $$F^{-1}(x) = \sum_{n=1}^\infty b_n \frac{x^n}{n!}$$ where $$b_n = \sum_{T \in S(\bar{L},n)} (-1)^{m(T)} w(T)$$ is the sum of the weights of <em>all</em> phylogenetic trees (since $\bar{L}$ is the set of all links.)</p> <p>See also <a href="https://digital.lib.washington.edu/researchworks/handle/1773/36757" rel="nofollow noreferrer">my thesis</a>, which contains a different generalization of your result. The phylogenetic trees (also called total partitions) are mentioned in Section 3.2</p> Answer 2: <p>This result can also be found as Theorem 3.3.9 in Ginzburg-Kapranov: "Koszul duality for operads" (Duke Math J 1994). Set $r=1$ in their result and rescale so that $a_1=1$. They had discovered it independently, but thank D. Wright for informing them that the result was known before due to J. Towber. They also give a reference to Wright: "The tree formulas for reversion of power series" (JPAA 1989), where this result appears as Theorem 3.10.</p> <p>The context of this result in the setting of operads is that any dg operad $P = \{P(n)\}$, say with $P(0)=0$ and $P(1)$ a copy of the base field, has a generating series which encodes the Euler characteristics of the chain complexes $P(n)$. The generating series of $P$ is the compositional inverse of the generating series for the "bar construction" on the operad $P$. The bar construction of operads is defined by a sum over trees whose internal vertices are decorated by the chain complexes of the operad $P$; in fact, the trees involved are exactly those that you call phylogenetic trees.</p> Answer 3: <p>See these three papers and references therein:</p> <p><a href="https://arxiv.org/abs/math/0208174" rel="nofollow noreferrer">https://arxiv.org/abs/math/0208174</a></p> <p><a href="https://arxiv.org/abs/math/0208173" rel="nofollow noreferrer">https://arxiv.org/abs/math/0208173</a></p> <p><a href="http://www.emis.de/journals/SLC/wpapers/s49abdess.html" rel="nofollow noreferrer">http://www.emis.de/journals/SLC/wpapers/s49abdess.html</a></p> <p>Going from one type of tree to another is no big deal in the framework of Joyal's theory of combinatorial species. Types of trees correspond to certain functors between finite sets and for each such functors there is an associated exponential generating series. Changing from say planar trees to Cayley trees is a morphism of functors and there is a precise theorem which says what happens then to the generating series (Theorem 8 in the last paper I listed).</p> Answer 4: <p>Yes, this is known. As Tom Copeland mentioned, this is in <a href="http://people.brandeis.edu/~gessel/homepage/students/drakethesis.pdf" rel="nofollow noreferrer">Drake's thesis</a>, Example 1.4.7.</p> <p>Drake gives a broad generalization of your result. His main theorem (Theorem 1.3.3) is roughly as follows. Suppose you have an alphabet $A$ of certain trees, and a set $L$ of 'allowed links' between them, ways of connecting the root of one to a leaf of another. Let $S(L,n)$ be the set of all trees $T$ with $n$ leaves labeled bijectively $1, \ldots, n$ with no other nodes labeled with the property that $T$ made up of the letters $A$ with <em>only</em> links between them which are in $L$. Say each tree $T \in A$ has a weight $w(T)$, and a tree made up of letters $T_1, \ldots, T_n \in A$ with certain links between them has weight $w(T_1) \cdots w(T_n)$. Let $$F(x) = \sum_{n=1}^\infty a_n \frac{x^n}{n!}$$ where $a_n = \sum_{T \in S(L,n)} w(T)$. Then the compositional inverse is $$F^{-1}(x) = \sum_{n=1}^\infty b_n \frac{x^n}{n!}$$ where $$b_n = \sum_{T \in S(\bar{L},n)} (-1)^{m(T)} w(T).$$ Here $m(T)$ is the number of letters in $A$ that $T$ is composed of, and $\bar{L}$ is the complement of $L$ in the set of all possible links between elements of $A$.</p> <p>This was an extension of the work of Susan Parker, who had a similar theorem using ordinary generating functions. Parker's result can, in turn, be seen as a generalization of a result of Carlitz, Scoville and Vaughan. Ira Gessel's <a href="http://people.brandeis.edu/~gessel/homepage/slides/csv.pdf" rel="nofollow noreferrer">slides</a> may be helpful.</p> <p>To get your theorem, let $A$ be the set of "simple" trees, the trees $T_n$, $n &gt; 1$, so that $T_n$ is a tree with the root having the $n$ leaves $1,2, \ldots, n$ as children. Let $w(T_n) = t_n$ (an indeterminate). Let $L$ be the empty set of links. Then $$F(x) = x + t_2\frac{x^2}{2!} + t_3 \frac{x^3}{3!} + \cdots$$ since $S(L,n) = \{T_n\}$ for $n&gt;1$, and $S(L,1)$ consists of the tree with one node which is always allowed (and has weight $1$).</p> <p>Then by Drake's inversion theorem, $$F^{-1}(x) = \sum_{n=1}^\infty b_n \frac{x^n}{n!}$$ where $$b_n = \sum_{T \in S(\bar{L},n)} (-1)^{m(T)} w(T)$$ is the sum of the weights of <em>all</em> phylogenetic trees (since $\bar{L}$ is the set of all links.)</p> <p>See also <a href="https://digital.lib.washington.edu/researchworks/handle/1773/36757" rel="nofollow noreferrer">my thesis</a>, which contains a different generalization of your result. The phylogenetic trees (also called total partitions) are mentioned in Section 3.2</p> Answer 5: <p>This result can also be found as Theorem 3.3.9 in Ginzburg-Kapranov: "Koszul duality for operads" (Duke Math J 1994). Set $r=1$ in their result and rescale so that $a_1=1$. They had discovered it independently, but thank D. Wright for informing them that the result was known before due to J. Towber. They also give a reference to Wright: "The tree formulas for reversion of power series" (JPAA 1989), where this result appears as Theorem 3.10.</p> <p>The context of this result in the setting of operads is that any dg operad $P = \{P(n)\}$, say with $P(0)=0$ and $P(1)$ a copy of the base field, has a generating series which encodes the Euler characteristics of the chain complexes $P(n)$. The generating series of $P$ is the compositional inverse of the generating series for the "bar construction" on the operad $P$. The bar construction of operads is defined by a sum over trees whose internal vertices are decorated by the chain complexes of the operad $P$; in fact, the trees involved are exactly those that you call phylogenetic trees.</p> Answer 6: <p>See these three papers and references therein:</p> <p><a href="https://arxiv.org/abs/math/0208174" rel="nofollow noreferrer">https://arxiv.org/abs/math/0208174</a></p> <p><a href="https://arxiv.org/abs/math/0208173" rel="nofollow noreferrer">https://arxiv.org/abs/math/0208173</a></p> <p><a href="http://www.emis.de/journals/SLC/wpapers/s49abdess.html" rel="nofollow noreferrer">http://www.emis.de/journals/SLC/wpapers/s49abdess.html</a></p> <p>Going from one type of tree to another is no big deal in the framework of Joyal's theory of combinatorial species. Types of trees correspond to certain functors between finite sets and for each such functors there is an associated exponential generating series. Changing from say planar trees to Cayley trees is a morphism of functors and there is a precise theorem which says what happens then to the generating series (Theorem 8 in the last paper I listed).</p>
Irreducible representations of simple complex groups
259,491
8
https://mathoverflow.net/questions/259491/irreducible-representations-of-simple-complex-groups
<p>Let $G$ be a simple complex algebraic group. What are its complex irreducible finite-dimensional representations?</p> <p>Before you start voting to close the question, I never said "rational". I am asking about abstract representations of $G$ as a group.</p> <p>I can a make an obvious conjecture but I have no idea how to prove it, even for $SL_2$. Yet, it could be something well-known. Pick a rational irreducible representation $(V,\rho)$ and a field endomorphism $\tau: {\mathbb C} \rightarrow {\mathbb C}$. The field endomorphism extends to an endmorphism of $G$: just apply $\tau$ to each matrix entry of of $g\in G\subseteq GL_n(\mathbb C)$. Thus, we get a twisted rational representation $(V,\tau\circ \rho)$ that clearly remains irreducible.</p> <p>STUPID WABBIT CONJECTURE: all irreducible representations are twisted rational representations. </p>
[ { "answer_id": 259512, "score": 17, "text": "<p>If $\\rho:G(\\mathbf{C}) \\rightarrow {\\rm{GL}}(V)(\\mathbf{C})$ is such an abstract linear representation, by irreducibility (or mere semisimplicity) the Zariski closure $H \\subset {\\rm{GL}}(V)$ of the image has connected reductive identity component (due to Lie-Kolchin). But $G(\\mathbf{C})$ is perfect with simple quotient modulo its finite center, so $H^0$ is semisimple, and then even simple in the sense of algebraic groups. The group $G(\\mathbf{C})$ is generated by its \"unipotent\" subgroups that have no nontrivial finite quotient, so $H$ is even connected.</p>\n\n<p>Thus, your question amounts to asking if $G(\\mathbf{C}) \\rightarrow H(\\mathbf{C})$ arises from a $\\mathbf{C}$-homomorphism $G \\rightarrow H$ up to a field map $\\mathbf{C} \\rightarrow \\mathbf{C}$. This is a (very) special case of the setting that was completely analyzed in an optimal way by Borel and Tits in their paper <em>Homomorphismes \"abstraits\" de groupes algébriques simples</em> in Annals of Math 97 (1973), pp. 499-571. There they work over general infinite fields, with connnected semisimple groups that are absolutely simple. The Introduction presents their main results (A) and (B) which address your question (in view of the initial part of this answer). See 8.18 for a certain standard counterexample over $\\mathbf{R}$ and 8.19 for a conjecture in wider generality which requires some caveats in characteristic 2 (the necessity of which is explained in a Remark in the Introduction of Igor Rapinchuk's 2013 paper \"On abstract representations of the groups of rational points of algebraic groups and their deformations\" in Algebra and Number Theory, vol. 7). </p>\n" }, { "answer_id": 262613, "score": 5, "text": "<p>Just to complete the answer, every complex irreducible representation of $G(\\mathbb{C})$ ($G$ is simply connected) is an $r$-fold tensor product of the form $\\rho _1\\circ \\sigma _1\\otimes \\cdots \\otimes \\rho _r \\circ \\sigma _r$, where each $\\rho _i$ is an algebraic irreducible representation of $G$ and $\\sigma _i$ are $distinct$ embeddings of the field $\\mathbb{C}$ into itself. This readily follows from the results of Borel-Tits (see Theorem (10.3) of the article of Borel-Tits mentioned by nfdc23). </p>\n" }, { "answer_id": 259512, "score": 17, "text": "<p>If $\\rho:G(\\mathbf{C}) \\rightarrow {\\rm{GL}}(V)(\\mathbf{C})$ is such an abstract linear representation, by irreducibility (or mere semisimplicity) the Zariski closure $H \\subset {\\rm{GL}}(V)$ of the image has connected reductive identity component (due to Lie-Kolchin). But $G(\\mathbf{C})$ is perfect with simple quotient modulo its finite center, so $H^0$ is semisimple, and then even simple in the sense of algebraic groups. The group $G(\\mathbf{C})$ is generated by its \"unipotent\" subgroups that have no nontrivial finite quotient, so $H$ is even connected.</p>\n\n<p>Thus, your question amounts to asking if $G(\\mathbf{C}) \\rightarrow H(\\mathbf{C})$ arises from a $\\mathbf{C}$-homomorphism $G \\rightarrow H$ up to a field map $\\mathbf{C} \\rightarrow \\mathbf{C}$. This is a (very) special case of the setting that was completely analyzed in an optimal way by Borel and Tits in their paper <em>Homomorphismes \"abstraits\" de groupes algébriques simples</em> in Annals of Math 97 (1973), pp. 499-571. There they work over general infinite fields, with connnected semisimple groups that are absolutely simple. The Introduction presents their main results (A) and (B) which address your question (in view of the initial part of this answer). See 8.18 for a certain standard counterexample over $\\mathbf{R}$ and 8.19 for a conjecture in wider generality which requires some caveats in characteristic 2 (the necessity of which is explained in a Remark in the Introduction of Igor Rapinchuk's 2013 paper \"On abstract representations of the groups of rational points of algebraic groups and their deformations\" in Algebra and Number Theory, vol. 7). </p>\n" }, { "answer_id": 262613, "score": 5, "text": "<p>Just to complete the answer, every complex irreducible representation of $G(\\mathbb{C})$ ($G$ is simply connected) is an $r$-fold tensor product of the form $\\rho _1\\circ \\sigma _1\\otimes \\cdots \\otimes \\rho _r \\circ \\sigma _r$, where each $\\rho _i$ is an algebraic irreducible representation of $G$ and $\\sigma _i$ are $distinct$ embeddings of the field $\\mathbb{C}$ into itself. This readily follows from the results of Borel-Tits (see Theorem (10.3) of the article of Borel-Tits mentioned by nfdc23). </p>\n" } ]
Question: Irreducible representations of simple complex groups Answer 1: <p>If $\rho:G(\mathbf{C}) \rightarrow {\rm{GL}}(V)(\mathbf{C})$ is such an abstract linear representation, by irreducibility (or mere semisimplicity) the Zariski closure $H \subset {\rm{GL}}(V)$ of the image has connected reductive identity component (due to Lie-Kolchin). But $G(\mathbf{C})$ is perfect with simple quotient modulo its finite center, so $H^0$ is semisimple, and then even simple in the sense of algebraic groups. The group $G(\mathbf{C})$ is generated by its "unipotent" subgroups that have no nontrivial finite quotient, so $H$ is even connected.</p> <p>Thus, your question amounts to asking if $G(\mathbf{C}) \rightarrow H(\mathbf{C})$ arises from a $\mathbf{C}$-homomorphism $G \rightarrow H$ up to a field map $\mathbf{C} \rightarrow \mathbf{C}$. This is a (very) special case of the setting that was completely analyzed in an optimal way by Borel and Tits in their paper <em>Homomorphismes "abstraits" de groupes algébriques simples</em> in Annals of Math 97 (1973), pp. 499-571. There they work over general infinite fields, with connnected semisimple groups that are absolutely simple. The Introduction presents their main results (A) and (B) which address your question (in view of the initial part of this answer). See 8.18 for a certain standard counterexample over $\mathbf{R}$ and 8.19 for a conjecture in wider generality which requires some caveats in characteristic 2 (the necessity of which is explained in a Remark in the Introduction of Igor Rapinchuk's 2013 paper "On abstract representations of the groups of rational points of algebraic groups and their deformations" in Algebra and Number Theory, vol. 7). </p> Answer 2: <p>Just to complete the answer, every complex irreducible representation of $G(\mathbb{C})$ ($G$ is simply connected) is an $r$-fold tensor product of the form $\rho _1\circ \sigma _1\otimes \cdots \otimes \rho _r \circ \sigma _r$, where each $\rho _i$ is an algebraic irreducible representation of $G$ and $\sigma _i$ are $distinct$ embeddings of the field $\mathbb{C}$ into itself. This readily follows from the results of Borel-Tits (see Theorem (10.3) of the article of Borel-Tits mentioned by nfdc23). </p> Answer 3: <p>If $\rho:G(\mathbf{C}) \rightarrow {\rm{GL}}(V)(\mathbf{C})$ is such an abstract linear representation, by irreducibility (or mere semisimplicity) the Zariski closure $H \subset {\rm{GL}}(V)$ of the image has connected reductive identity component (due to Lie-Kolchin). But $G(\mathbf{C})$ is perfect with simple quotient modulo its finite center, so $H^0$ is semisimple, and then even simple in the sense of algebraic groups. The group $G(\mathbf{C})$ is generated by its "unipotent" subgroups that have no nontrivial finite quotient, so $H$ is even connected.</p> <p>Thus, your question amounts to asking if $G(\mathbf{C}) \rightarrow H(\mathbf{C})$ arises from a $\mathbf{C}$-homomorphism $G \rightarrow H$ up to a field map $\mathbf{C} \rightarrow \mathbf{C}$. This is a (very) special case of the setting that was completely analyzed in an optimal way by Borel and Tits in their paper <em>Homomorphismes "abstraits" de groupes algébriques simples</em> in Annals of Math 97 (1973), pp. 499-571. There they work over general infinite fields, with connnected semisimple groups that are absolutely simple. The Introduction presents their main results (A) and (B) which address your question (in view of the initial part of this answer). See 8.18 for a certain standard counterexample over $\mathbf{R}$ and 8.19 for a conjecture in wider generality which requires some caveats in characteristic 2 (the necessity of which is explained in a Remark in the Introduction of Igor Rapinchuk's 2013 paper "On abstract representations of the groups of rational points of algebraic groups and their deformations" in Algebra and Number Theory, vol. 7). </p> Answer 4: <p>Just to complete the answer, every complex irreducible representation of $G(\mathbb{C})$ ($G$ is simply connected) is an $r$-fold tensor product of the form $\rho _1\circ \sigma _1\otimes \cdots \otimes \rho _r \circ \sigma _r$, where each $\rho _i$ is an algebraic irreducible representation of $G$ and $\sigma _i$ are $distinct$ embeddings of the field $\mathbb{C}$ into itself. This readily follows from the results of Borel-Tits (see Theorem (10.3) of the article of Borel-Tits mentioned by nfdc23). </p>
Proof of Karlin-Rubin&#39;s theorem
259,598
8
https://mathoverflow.net/questions/259598/proof-of-karlin-rubins-theorem
<p>I asked this question on Math Exchange, but as I did not receive a successful answer, maybe you could help me.</p> <p>Karlin-Rubin's theorem states conditions under which we can find a uniformly most powerful test (UMPT) for a statistical hypothesis:</p> <blockquote> <p>Suppose a family of density or mass functions $\{f(\vec{x}|\theta):\,\theta\in\Theta\}$ and we want to test $$\begin{cases} H_0:\,\theta\leq\theta_0 \\ H_A:\,\theta&gt;\theta_0.\end{cases}$$If the likelihood ratio is monotone on a statistic $T(\vec{x})$ (that is, for every fixed $\theta_1&lt;\theta_2$ in $\Theta$, the ratio $\frac{f(\vec{x}|\theta_2)}{f(\vec{x}|\theta_1)}$ is nondecreasing on $\{\vec{x}:\,f(\vec{x}|\theta_2)&gt;0\text{ or }f(\vec{x}|\theta_1)&gt;0\}$ as a function of $T(\vec{x})$, interpreting $c/0=\infty$ if $c&gt;0$), then the test of critical region $\text{CR}=\{\vec{x}:\,T(\vec{x})\geq k\}$, where $k$ is chosen so that $\alpha=P(\text{CR}|\theta=\theta_0)$, is the UMPT of size $\alpha$.</p> </blockquote> <p>In all the proofs I have read (for instance, in page 22 <a href="https://www.soph.uab.edu/sites/edu.ssg/files/People/KZhang/BST695-11Fall/BST695-2011-Fall-Chapter-08.pdf" rel="noreferrer">here</a> or in "<a href="http://people.unica.it/musio/files/2008/10/Casella-Berger.pdf" rel="noreferrer">Statistical inference</a>" by Casella-Berger, 2n edition, page 391), it is (more or less) said: "we can find $k_1$ such that, if $T(\vec{x})\geq k$, then $\frac{f(\vec{x}|\theta_2)}{f(\vec{x}|\theta_1)}\geq k_1$, and if $T(\vec{x})&lt;k$, then $\frac{f(\vec{x}|\theta_2)}{f(\vec{x}|\theta_1)}&lt; k_1$". I would understand that statement if the likehood ratio were strictly increasing, but what about the case in which it is constant? </p> <p>For example, if $X\sim U(0,\theta)$, the likelihood ratio is monotone on $T(\vec{x})=\max_{1\leq i\leq n}x_i$ ($n$ is the length of the sample $\vec{x}$), but not strictly increasing. </p> <p>My questions are: </p> <ol> <li><p>Is the assertion between quotation marks true for every density or mass function with (not strictly) monotone likelihood ratio on $T$? </p></li> <li><p>And what about in the case of the uniform distribution?</p></li> </ol>
[ { "answer_id": 377514, "score": 2, "text": "<p><span class=\"math-container\">$\\newcommand\\th\\theta\\newcommand\\al\\alpha$</span>Even the following more general assertion is true:</p>\n<blockquote>\n<p>Suppose that there is a statistic <span class=\"math-container\">$T$</span> such that for all <span class=\"math-container\">$\\th_0$</span> and <span class=\"math-container\">$\\th_1$</span> such that <span class=\"math-container\">$\\th_0&lt;\\th_1$</span> there is a nondecreasing function <span class=\"math-container\">$g_{\\th_0,\\th_1}$</span> such that for all <span class=\"math-container\">$x$</span> we have\n<span class=\"math-container\">$$r_{\\th_0,\\th_1}(x):=\\frac{f_{\\th_1}(x)}{f_{\\th_0}(x)}=g_{\\th_0,\\th_1}(T(x)).$$</span>\nSuppose that a test <span class=\"math-container\">$\\phi$</span> is such that for some real <span class=\"math-container\">$k$</span> and all <span class=\"math-container\">$x$</span>\n<span class=\"math-container\">$$\\phi(x)=\\begin{cases}\n1&amp;\\text{ if }T(x)&gt;k, \\\\ \n0&amp;\\text{ if }T(x)&lt;k \n\\end{cases}$$</span>\n(the test <span class=\"math-container\">$\\phi$</span> may be randomized, taking any value <span class=\"math-container\">$\\phi(x)$</span> in the interval <span class=\"math-container\">$[0,1]$</span> if <span class=\"math-container\">$x$</span> is such that <span class=\"math-container\">$T(x)=k$</span>).\nThen the test <span class=\"math-container\">$\\phi$</span> is uniformly most powerful of level <span class=\"math-container\">$\\al:=E\\phi(X)$</span> for the hypotheses as in the OP.</p>\n</blockquote>\n<p>This follows because, for <span class=\"math-container\">$c:=g_{\\th_0,\\th_1}(k)$</span>, we have the implications <span class=\"math-container\">$$r_{\\th_0,\\th_1}(x)&gt;c\\iff g_{\\th_0,\\th_1}(T(x))&gt;c\\implies T(x)&gt;k\\implies\\phi(x)=1\\tag{1}$$</span>\n(so that <span class=\"math-container\">$H_0$</span> is rejected by the test <span class=\"math-container\">$\\phi$</span>) and\n<span class=\"math-container\">$$r_{\\th_0,\\th_1}(x)&lt;c\\iff g_{\\th_0,\\th_1}(T(x))&lt;c\\implies T(x)&lt;k\\implies\\phi(x)=0\\tag{2}$$</span>\n(so that <span class=\"math-container\">$H_0$</span> is not rejected by the test <span class=\"math-container\">$\\phi$</span>) -- which means that <span class=\"math-container\">$\\phi$</span> is a <a href=\"https://web.stanford.edu/%7Elmackey/stats300a/doc/stats300a-fall15-lecture13.pdf\" rel=\"nofollow noreferrer\">Neyman--Pearson test</a>for any <span class=\"math-container\">$\\th_0$</span> and <span class=\"math-container\">$\\th_1$</span> such that <span class=\"math-container\">$\\th_0&lt;\\th_1$</span>.</p>\n<p>That <span class=\"math-container\">$T$</span> is not strictly increasing causes no problems whatsoever, because in (1) and (2) we only need the left-to-right implications.</p>\n" }, { "answer_id": 259647, "score": 0, "text": "<p>If it were to be constant then the statistical model $\\{f(\\vec{x}|\\theta):\\,\\theta\\in\\Theta\\}$ would only contain copies of the same density, i.e. one single density. In this case, inference is quite useless, but nonetheless the test is UMPT (and unbiased) since power and size agree for every test.</p>\n" }, { "answer_id": 377514, "score": 2, "text": "<p><span class=\"math-container\">$\\newcommand\\th\\theta\\newcommand\\al\\alpha$</span>Even the following more general assertion is true:</p>\n<blockquote>\n<p>Suppose that there is a statistic <span class=\"math-container\">$T$</span> such that for all <span class=\"math-container\">$\\th_0$</span> and <span class=\"math-container\">$\\th_1$</span> such that <span class=\"math-container\">$\\th_0&lt;\\th_1$</span> there is a nondecreasing function <span class=\"math-container\">$g_{\\th_0,\\th_1}$</span> such that for all <span class=\"math-container\">$x$</span> we have\n<span class=\"math-container\">$$r_{\\th_0,\\th_1}(x):=\\frac{f_{\\th_1}(x)}{f_{\\th_0}(x)}=g_{\\th_0,\\th_1}(T(x)).$$</span>\nSuppose that a test <span class=\"math-container\">$\\phi$</span> is such that for some real <span class=\"math-container\">$k$</span> and all <span class=\"math-container\">$x$</span>\n<span class=\"math-container\">$$\\phi(x)=\\begin{cases}\n1&amp;\\text{ if }T(x)&gt;k, \\\\ \n0&amp;\\text{ if }T(x)&lt;k \n\\end{cases}$$</span>\n(the test <span class=\"math-container\">$\\phi$</span> may be randomized, taking any value <span class=\"math-container\">$\\phi(x)$</span> in the interval <span class=\"math-container\">$[0,1]$</span> if <span class=\"math-container\">$x$</span> is such that <span class=\"math-container\">$T(x)=k$</span>).\nThen the test <span class=\"math-container\">$\\phi$</span> is uniformly most powerful of level <span class=\"math-container\">$\\al:=E\\phi(X)$</span> for the hypotheses as in the OP.</p>\n</blockquote>\n<p>This follows because, for <span class=\"math-container\">$c:=g_{\\th_0,\\th_1}(k)$</span>, we have the implications <span class=\"math-container\">$$r_{\\th_0,\\th_1}(x)&gt;c\\iff g_{\\th_0,\\th_1}(T(x))&gt;c\\implies T(x)&gt;k\\implies\\phi(x)=1\\tag{1}$$</span>\n(so that <span class=\"math-container\">$H_0$</span> is rejected by the test <span class=\"math-container\">$\\phi$</span>) and\n<span class=\"math-container\">$$r_{\\th_0,\\th_1}(x)&lt;c\\iff g_{\\th_0,\\th_1}(T(x))&lt;c\\implies T(x)&lt;k\\implies\\phi(x)=0\\tag{2}$$</span>\n(so that <span class=\"math-container\">$H_0$</span> is not rejected by the test <span class=\"math-container\">$\\phi$</span>) -- which means that <span class=\"math-container\">$\\phi$</span> is a <a href=\"https://web.stanford.edu/%7Elmackey/stats300a/doc/stats300a-fall15-lecture13.pdf\" rel=\"nofollow noreferrer\">Neyman--Pearson test</a>for any <span class=\"math-container\">$\\th_0$</span> and <span class=\"math-container\">$\\th_1$</span> such that <span class=\"math-container\">$\\th_0&lt;\\th_1$</span>.</p>\n<p>That <span class=\"math-container\">$T$</span> is not strictly increasing causes no problems whatsoever, because in (1) and (2) we only need the left-to-right implications.</p>\n" }, { "answer_id": 259647, "score": 0, "text": "<p>If it were to be constant then the statistical model $\\{f(\\vec{x}|\\theta):\\,\\theta\\in\\Theta\\}$ would only contain copies of the same density, i.e. one single density. In this case, inference is quite useless, but nonetheless the test is UMPT (and unbiased) since power and size agree for every test.</p>\n" } ]
Question: Proof of Karlin-Rubin&#39;s theorem Answer 1: <p><span class="math-container">$\newcommand\th\theta\newcommand\al\alpha$</span>Even the following more general assertion is true:</p> <blockquote> <p>Suppose that there is a statistic <span class="math-container">$T$</span> such that for all <span class="math-container">$\th_0$</span> and <span class="math-container">$\th_1$</span> such that <span class="math-container">$\th_0&lt;\th_1$</span> there is a nondecreasing function <span class="math-container">$g_{\th_0,\th_1}$</span> such that for all <span class="math-container">$x$</span> we have <span class="math-container">$$r_{\th_0,\th_1}(x):=\frac{f_{\th_1}(x)}{f_{\th_0}(x)}=g_{\th_0,\th_1}(T(x)).$$</span> Suppose that a test <span class="math-container">$\phi$</span> is such that for some real <span class="math-container">$k$</span> and all <span class="math-container">$x$</span> <span class="math-container">$$\phi(x)=\begin{cases} 1&amp;\text{ if }T(x)&gt;k, \\ 0&amp;\text{ if }T(x)&lt;k \end{cases}$$</span> (the test <span class="math-container">$\phi$</span> may be randomized, taking any value <span class="math-container">$\phi(x)$</span> in the interval <span class="math-container">$[0,1]$</span> if <span class="math-container">$x$</span> is such that <span class="math-container">$T(x)=k$</span>). Then the test <span class="math-container">$\phi$</span> is uniformly most powerful of level <span class="math-container">$\al:=E\phi(X)$</span> for the hypotheses as in the OP.</p> </blockquote> <p>This follows because, for <span class="math-container">$c:=g_{\th_0,\th_1}(k)$</span>, we have the implications <span class="math-container">$$r_{\th_0,\th_1}(x)&gt;c\iff g_{\th_0,\th_1}(T(x))&gt;c\implies T(x)&gt;k\implies\phi(x)=1\tag{1}$$</span> (so that <span class="math-container">$H_0$</span> is rejected by the test <span class="math-container">$\phi$</span>) and <span class="math-container">$$r_{\th_0,\th_1}(x)&lt;c\iff g_{\th_0,\th_1}(T(x))&lt;c\implies T(x)&lt;k\implies\phi(x)=0\tag{2}$$</span> (so that <span class="math-container">$H_0$</span> is not rejected by the test <span class="math-container">$\phi$</span>) -- which means that <span class="math-container">$\phi$</span> is a <a href="https://web.stanford.edu/%7Elmackey/stats300a/doc/stats300a-fall15-lecture13.pdf" rel="nofollow noreferrer">Neyman--Pearson test</a>for any <span class="math-container">$\th_0$</span> and <span class="math-container">$\th_1$</span> such that <span class="math-container">$\th_0&lt;\th_1$</span>.</p> <p>That <span class="math-container">$T$</span> is not strictly increasing causes no problems whatsoever, because in (1) and (2) we only need the left-to-right implications.</p> Answer 2: <p>If it were to be constant then the statistical model $\{f(\vec{x}|\theta):\,\theta\in\Theta\}$ would only contain copies of the same density, i.e. one single density. In this case, inference is quite useless, but nonetheless the test is UMPT (and unbiased) since power and size agree for every test.</p> Answer 3: <p><span class="math-container">$\newcommand\th\theta\newcommand\al\alpha$</span>Even the following more general assertion is true:</p> <blockquote> <p>Suppose that there is a statistic <span class="math-container">$T$</span> such that for all <span class="math-container">$\th_0$</span> and <span class="math-container">$\th_1$</span> such that <span class="math-container">$\th_0&lt;\th_1$</span> there is a nondecreasing function <span class="math-container">$g_{\th_0,\th_1}$</span> such that for all <span class="math-container">$x$</span> we have <span class="math-container">$$r_{\th_0,\th_1}(x):=\frac{f_{\th_1}(x)}{f_{\th_0}(x)}=g_{\th_0,\th_1}(T(x)).$$</span> Suppose that a test <span class="math-container">$\phi$</span> is such that for some real <span class="math-container">$k$</span> and all <span class="math-container">$x$</span> <span class="math-container">$$\phi(x)=\begin{cases} 1&amp;\text{ if }T(x)&gt;k, \\ 0&amp;\text{ if }T(x)&lt;k \end{cases}$$</span> (the test <span class="math-container">$\phi$</span> may be randomized, taking any value <span class="math-container">$\phi(x)$</span> in the interval <span class="math-container">$[0,1]$</span> if <span class="math-container">$x$</span> is such that <span class="math-container">$T(x)=k$</span>). Then the test <span class="math-container">$\phi$</span> is uniformly most powerful of level <span class="math-container">$\al:=E\phi(X)$</span> for the hypotheses as in the OP.</p> </blockquote> <p>This follows because, for <span class="math-container">$c:=g_{\th_0,\th_1}(k)$</span>, we have the implications <span class="math-container">$$r_{\th_0,\th_1}(x)&gt;c\iff g_{\th_0,\th_1}(T(x))&gt;c\implies T(x)&gt;k\implies\phi(x)=1\tag{1}$$</span> (so that <span class="math-container">$H_0$</span> is rejected by the test <span class="math-container">$\phi$</span>) and <span class="math-container">$$r_{\th_0,\th_1}(x)&lt;c\iff g_{\th_0,\th_1}(T(x))&lt;c\implies T(x)&lt;k\implies\phi(x)=0\tag{2}$$</span> (so that <span class="math-container">$H_0$</span> is not rejected by the test <span class="math-container">$\phi$</span>) -- which means that <span class="math-container">$\phi$</span> is a <a href="https://web.stanford.edu/%7Elmackey/stats300a/doc/stats300a-fall15-lecture13.pdf" rel="nofollow noreferrer">Neyman--Pearson test</a>for any <span class="math-container">$\th_0$</span> and <span class="math-container">$\th_1$</span> such that <span class="math-container">$\th_0&lt;\th_1$</span>.</p> <p>That <span class="math-container">$T$</span> is not strictly increasing causes no problems whatsoever, because in (1) and (2) we only need the left-to-right implications.</p> Answer 4: <p>If it were to be constant then the statistical model $\{f(\vec{x}|\theta):\,\theta\in\Theta\}$ would only contain copies of the same density, i.e. one single density. In this case, inference is quite useless, but nonetheless the test is UMPT (and unbiased) since power and size agree for every test.</p>
Is there a model structure on (strict?) Monoidal Categories?
259,734
8
https://mathoverflow.net/questions/259734/is-there-a-model-structure-on-strict-monoidal-categories
<p>Basically, what the title says. </p> <p>Presumably, one could use the fact that monoidal categories (resp. strict monoidal categories) are one-object bicategories (resp. 2-categories) and use the Lack model structure on those, but I am unsure if this would work or not.</p>
[ { "answer_id": 259737, "score": 9, "text": "<p>I'm not sure about the case of general monoidal categories. (Although I seem to recall a remark that there is no such structure since the category of monoidal categories is not cocomplete and a suitable replacement would be the category of multicategories. Perhaps somebody can confirm this.)</p>\n\n<p>However, the case of strict monoidal categories is known since they are simply monoids internal to the category of categories. It is a theorem of Schwede and Shipley (Theorem 4.1 in <em>Algebras and Modules in Monoidal Model Categories</em>) that if a monoidal model category satisfies so called <em>monoid axiom</em>, then the category of monoids in it inherits a model structure. The monoid axiom is a technical condition which is, in particular, automatically satisfied if all objects are cofibrant as is the case in the canonical model structure on the category of categories.</p>\n" }, { "answer_id": 259787, "score": 4, "text": "<p>There is a model structure on the category of monoidal categories and strict monoidal functors. A strict monoidal functor is a fibration / weak equivalence just when its underlying functor is a fibration / weak equivalence in Cat.</p>\n\n<p>More generally for T a 2-monad with rank on Cat you can lift the model structure on Cat to the category of strict T-algebras and strict T-algebra morphisms: this generalises the above example. </p>\n\n<p>For these results, see the discussion in Section 1.7 and Theorem 4.5 of Steve Lack's paper \"Homotopy theoretic aspects of 2-monads\": <a href=\"https://arxiv.org/abs/math/0607646\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/math/0607646</a></p>\n" }, { "answer_id": 259737, "score": 9, "text": "<p>I'm not sure about the case of general monoidal categories. (Although I seem to recall a remark that there is no such structure since the category of monoidal categories is not cocomplete and a suitable replacement would be the category of multicategories. Perhaps somebody can confirm this.)</p>\n\n<p>However, the case of strict monoidal categories is known since they are simply monoids internal to the category of categories. It is a theorem of Schwede and Shipley (Theorem 4.1 in <em>Algebras and Modules in Monoidal Model Categories</em>) that if a monoidal model category satisfies so called <em>monoid axiom</em>, then the category of monoids in it inherits a model structure. The monoid axiom is a technical condition which is, in particular, automatically satisfied if all objects are cofibrant as is the case in the canonical model structure on the category of categories.</p>\n" }, { "answer_id": 259787, "score": 4, "text": "<p>There is a model structure on the category of monoidal categories and strict monoidal functors. A strict monoidal functor is a fibration / weak equivalence just when its underlying functor is a fibration / weak equivalence in Cat.</p>\n\n<p>More generally for T a 2-monad with rank on Cat you can lift the model structure on Cat to the category of strict T-algebras and strict T-algebra morphisms: this generalises the above example. </p>\n\n<p>For these results, see the discussion in Section 1.7 and Theorem 4.5 of Steve Lack's paper \"Homotopy theoretic aspects of 2-monads\": <a href=\"https://arxiv.org/abs/math/0607646\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/math/0607646</a></p>\n" } ]
Question: Is there a model structure on (strict?) Monoidal Categories? Answer 1: <p>I'm not sure about the case of general monoidal categories. (Although I seem to recall a remark that there is no such structure since the category of monoidal categories is not cocomplete and a suitable replacement would be the category of multicategories. Perhaps somebody can confirm this.)</p> <p>However, the case of strict monoidal categories is known since they are simply monoids internal to the category of categories. It is a theorem of Schwede and Shipley (Theorem 4.1 in <em>Algebras and Modules in Monoidal Model Categories</em>) that if a monoidal model category satisfies so called <em>monoid axiom</em>, then the category of monoids in it inherits a model structure. The monoid axiom is a technical condition which is, in particular, automatically satisfied if all objects are cofibrant as is the case in the canonical model structure on the category of categories.</p> Answer 2: <p>There is a model structure on the category of monoidal categories and strict monoidal functors. A strict monoidal functor is a fibration / weak equivalence just when its underlying functor is a fibration / weak equivalence in Cat.</p> <p>More generally for T a 2-monad with rank on Cat you can lift the model structure on Cat to the category of strict T-algebras and strict T-algebra morphisms: this generalises the above example. </p> <p>For these results, see the discussion in Section 1.7 and Theorem 4.5 of Steve Lack's paper "Homotopy theoretic aspects of 2-monads": <a href="https://arxiv.org/abs/math/0607646" rel="nofollow noreferrer">https://arxiv.org/abs/math/0607646</a></p> Answer 3: <p>I'm not sure about the case of general monoidal categories. (Although I seem to recall a remark that there is no such structure since the category of monoidal categories is not cocomplete and a suitable replacement would be the category of multicategories. Perhaps somebody can confirm this.)</p> <p>However, the case of strict monoidal categories is known since they are simply monoids internal to the category of categories. It is a theorem of Schwede and Shipley (Theorem 4.1 in <em>Algebras and Modules in Monoidal Model Categories</em>) that if a monoidal model category satisfies so called <em>monoid axiom</em>, then the category of monoids in it inherits a model structure. The monoid axiom is a technical condition which is, in particular, automatically satisfied if all objects are cofibrant as is the case in the canonical model structure on the category of categories.</p> Answer 4: <p>There is a model structure on the category of monoidal categories and strict monoidal functors. A strict monoidal functor is a fibration / weak equivalence just when its underlying functor is a fibration / weak equivalence in Cat.</p> <p>More generally for T a 2-monad with rank on Cat you can lift the model structure on Cat to the category of strict T-algebras and strict T-algebra morphisms: this generalises the above example. </p> <p>For these results, see the discussion in Section 1.7 and Theorem 4.5 of Steve Lack's paper "Homotopy theoretic aspects of 2-monads": <a href="https://arxiv.org/abs/math/0607646" rel="nofollow noreferrer">https://arxiv.org/abs/math/0607646</a></p>
Einstein Manifolds in dimension five?
259,863
8
https://mathoverflow.net/questions/259863/einstein-manifolds-in-dimension-five
<p>My teacher of Riemannian Geometry told the class about the problem of finding Einstein Manifolds in 5 dimension. My question is, what is the difficulty of this problem? Possessing a first course in Riemmanian geometry can I understand this?</p> <p>Thanks</p>
[ { "answer_id": 259870, "score": 17, "text": "<p><strong>Answer for the compact homogeneous case:</strong> Due to Alekseevsky, Dotti and Ferraris (Homogeneous Ricci positive 5-manifolds, Pacific. J. Math, 175, 1-12, 1996) we know that in the non-symmetric case, the classification of 5-dimensional compact homogeneous Einstein spaces reduces to the classification of Einstein metrics on the space $M_{a, b}=SU(2)\\times SU(2)/S^{1}$, with a specific embedding of $S^{1}$. Here, $a\\geq b\\geq 0$, $a&gt; 0$ are integers with ${\\rm gcd}(a, b)&gt;1$. </p>\n\n<p>Wang and Ziller (Einstein metrics on principal torus bundles, J. Differ. Geom., 31, (1990) 215-248), proved that the spaces $M_{a, b}$ for $(a, b)\\neq (1, 1)$ admit exactly one (up to isometry) homogeneous Einstein metric and the same result was obtained indepedently by Rodionov. Wang and Ziller also proved that $M_{1, 1}$ admits at least one invariant Einstein metric. Note that the isotropy representation of $M_{1, 1}$ decomposes into three irreducible submodules, two of them are isomorphic to each other, and thus a $SU(2)^{2}$-invariant Riemannian metric is depending on four positive real parameters. Now, $M_{1, 1}$ admits a second homogeneous Einstein metric, which is isometric to a product of symmetric metrics on $S^{3}\\times S^{2}$. In fact, the space $M_{1, 1}$ can be viewed as the unit tangent bundle $T_1S^{3}=SO(4)/SO(2)$ of the 3-sphere $S^{3}$, and thus as the Stiefel manifold $V_{2}(\\mathbb{R}^{4})$ of 2-frames in $\\mathbb{R}^{4}$.\nNotice also that the dimension $n=5$ and examples like the 5-sphere $S^{5}$, the product $S^{3}\\times S^{2}$ and the space $M_{1, 1}$, are very important examples of 5-dimensional homogeneous or cohomogeneity-one Sasaki-Einstein manifolds. Finally, the sphere $S^{5}$ can be considered either as the irreducible symmetric space $SO(6)/SO(5)$, or as $SU(3)/SU(2)$ with a certain invariant metric. The isotropy representation of the second presentation decomposes into the sum of two irreducible summands. The proof of the fact that the canonical metric of $S^{5}=SU(3)/SU(2)$ is the unique $SU(3)$-invariant Einstein metric, was given by Jensen.</p>\n\n<p><strong>Difficulty of constructing general Einstein metrics:</strong> It is good to notice that Einstein metrics are relatively rare among Riemannian metrics. Although no topological obstructions are known for big dimensions, it is difficult to construct such metrics because the Einstein equation is a system of second order PDEs and there is no general method known. In recent decades, Einstein manifolds became the subject of many studies (both of mathematicians and physicists). Motivated from classical mechanics and the well known fact that symmetry considerations can simplify the study of geometric problems, the current strategy for constructing examples of Einstein metrics is to assume some symmetry condition for the metric tensor itself, that means that there is a Lie group which acts on the manifold in a smooth way by isometries. Roughly speaking, important progress has been done for homogeneous Riemannian $G$-manifolds and cohomogeneity-one Riemannian $G$-manifolds. The fist case arises from a transitive action of a Lie group $G$ on $(M, g)$, and the second one appears when the Lie group $G$ acts on $(M, g)$ with principal orbits of codimension one. Also, Einstein metrics with special holonomy is a famous topic (with a lot of interest in physics) and there are several results for this kind of Einstein metrics, e.g. Einstein-Sasakian manifolds, 6-dim nearly Kähler manifolds, 7-dim nearly parallel $G_2$-manifolds, etc, and all of them admit real Killing spinors. Also, Ricci-flatness is related with parallel spinors.</p>\n\n<p><strong>Finally:</strong> Possessing a first course in Riemmanian geometry and having some basic knowledge of representation theory... it is not difficult to prove for example that $SO(6)/SO(5)$ admits a unique $SO(6)$-invariant Einstein metric. But other problems related with Einstein metrics may need more expertise and studying. </p>\n" }, { "answer_id": 259872, "score": 5, "text": "<p>For what it is worth, there are many examples in <a href=\"https://arxiv.org/abs/1103.0817\" rel=\"noreferrer\">Dezhong Chen's paper.</a></p>\n" }, { "answer_id": 267008, "score": 4, "text": "<p>Simply connected 5-manifolds are classified by simple invariants, second homology and second Stiefel-Whitney class. So one can consider the existence problem, i.e. which 5-manifolds admit an Einstein metric. Many are know to have an Eintein metric. See the paper by János Kollár <a href=\"https://arxiv.org/abs/math/0408184\" rel=\"nofollow noreferrer\">\"Einstein metrics on 5-dimensional Seifert bundles\"</a>.</p>\n" }, { "answer_id": 259870, "score": 17, "text": "<p><strong>Answer for the compact homogeneous case:</strong> Due to Alekseevsky, Dotti and Ferraris (Homogeneous Ricci positive 5-manifolds, Pacific. J. Math, 175, 1-12, 1996) we know that in the non-symmetric case, the classification of 5-dimensional compact homogeneous Einstein spaces reduces to the classification of Einstein metrics on the space $M_{a, b}=SU(2)\\times SU(2)/S^{1}$, with a specific embedding of $S^{1}$. Here, $a\\geq b\\geq 0$, $a&gt; 0$ are integers with ${\\rm gcd}(a, b)&gt;1$. </p>\n\n<p>Wang and Ziller (Einstein metrics on principal torus bundles, J. Differ. Geom., 31, (1990) 215-248), proved that the spaces $M_{a, b}$ for $(a, b)\\neq (1, 1)$ admit exactly one (up to isometry) homogeneous Einstein metric and the same result was obtained indepedently by Rodionov. Wang and Ziller also proved that $M_{1, 1}$ admits at least one invariant Einstein metric. Note that the isotropy representation of $M_{1, 1}$ decomposes into three irreducible submodules, two of them are isomorphic to each other, and thus a $SU(2)^{2}$-invariant Riemannian metric is depending on four positive real parameters. Now, $M_{1, 1}$ admits a second homogeneous Einstein metric, which is isometric to a product of symmetric metrics on $S^{3}\\times S^{2}$. In fact, the space $M_{1, 1}$ can be viewed as the unit tangent bundle $T_1S^{3}=SO(4)/SO(2)$ of the 3-sphere $S^{3}$, and thus as the Stiefel manifold $V_{2}(\\mathbb{R}^{4})$ of 2-frames in $\\mathbb{R}^{4}$.\nNotice also that the dimension $n=5$ and examples like the 5-sphere $S^{5}$, the product $S^{3}\\times S^{2}$ and the space $M_{1, 1}$, are very important examples of 5-dimensional homogeneous or cohomogeneity-one Sasaki-Einstein manifolds. Finally, the sphere $S^{5}$ can be considered either as the irreducible symmetric space $SO(6)/SO(5)$, or as $SU(3)/SU(2)$ with a certain invariant metric. The isotropy representation of the second presentation decomposes into the sum of two irreducible summands. The proof of the fact that the canonical metric of $S^{5}=SU(3)/SU(2)$ is the unique $SU(3)$-invariant Einstein metric, was given by Jensen.</p>\n\n<p><strong>Difficulty of constructing general Einstein metrics:</strong> It is good to notice that Einstein metrics are relatively rare among Riemannian metrics. Although no topological obstructions are known for big dimensions, it is difficult to construct such metrics because the Einstein equation is a system of second order PDEs and there is no general method known. In recent decades, Einstein manifolds became the subject of many studies (both of mathematicians and physicists). Motivated from classical mechanics and the well known fact that symmetry considerations can simplify the study of geometric problems, the current strategy for constructing examples of Einstein metrics is to assume some symmetry condition for the metric tensor itself, that means that there is a Lie group which acts on the manifold in a smooth way by isometries. Roughly speaking, important progress has been done for homogeneous Riemannian $G$-manifolds and cohomogeneity-one Riemannian $G$-manifolds. The fist case arises from a transitive action of a Lie group $G$ on $(M, g)$, and the second one appears when the Lie group $G$ acts on $(M, g)$ with principal orbits of codimension one. Also, Einstein metrics with special holonomy is a famous topic (with a lot of interest in physics) and there are several results for this kind of Einstein metrics, e.g. Einstein-Sasakian manifolds, 6-dim nearly Kähler manifolds, 7-dim nearly parallel $G_2$-manifolds, etc, and all of them admit real Killing spinors. Also, Ricci-flatness is related with parallel spinors.</p>\n\n<p><strong>Finally:</strong> Possessing a first course in Riemmanian geometry and having some basic knowledge of representation theory... it is not difficult to prove for example that $SO(6)/SO(5)$ admits a unique $SO(6)$-invariant Einstein metric. But other problems related with Einstein metrics may need more expertise and studying. </p>\n" }, { "answer_id": 259872, "score": 5, "text": "<p>For what it is worth, there are many examples in <a href=\"https://arxiv.org/abs/1103.0817\" rel=\"noreferrer\">Dezhong Chen's paper.</a></p>\n" }, { "answer_id": 267008, "score": 4, "text": "<p>Simply connected 5-manifolds are classified by simple invariants, second homology and second Stiefel-Whitney class. So one can consider the existence problem, i.e. which 5-manifolds admit an Einstein metric. Many are know to have an Eintein metric. See the paper by János Kollár <a href=\"https://arxiv.org/abs/math/0408184\" rel=\"nofollow noreferrer\">\"Einstein metrics on 5-dimensional Seifert bundles\"</a>.</p>\n" } ]
Question: Einstein Manifolds in dimension five? Answer 1: <p><strong>Answer for the compact homogeneous case:</strong> Due to Alekseevsky, Dotti and Ferraris (Homogeneous Ricci positive 5-manifolds, Pacific. J. Math, 175, 1-12, 1996) we know that in the non-symmetric case, the classification of 5-dimensional compact homogeneous Einstein spaces reduces to the classification of Einstein metrics on the space $M_{a, b}=SU(2)\times SU(2)/S^{1}$, with a specific embedding of $S^{1}$. Here, $a\geq b\geq 0$, $a&gt; 0$ are integers with ${\rm gcd}(a, b)&gt;1$. </p> <p>Wang and Ziller (Einstein metrics on principal torus bundles, J. Differ. Geom., 31, (1990) 215-248), proved that the spaces $M_{a, b}$ for $(a, b)\neq (1, 1)$ admit exactly one (up to isometry) homogeneous Einstein metric and the same result was obtained indepedently by Rodionov. Wang and Ziller also proved that $M_{1, 1}$ admits at least one invariant Einstein metric. Note that the isotropy representation of $M_{1, 1}$ decomposes into three irreducible submodules, two of them are isomorphic to each other, and thus a $SU(2)^{2}$-invariant Riemannian metric is depending on four positive real parameters. Now, $M_{1, 1}$ admits a second homogeneous Einstein metric, which is isometric to a product of symmetric metrics on $S^{3}\times S^{2}$. In fact, the space $M_{1, 1}$ can be viewed as the unit tangent bundle $T_1S^{3}=SO(4)/SO(2)$ of the 3-sphere $S^{3}$, and thus as the Stiefel manifold $V_{2}(\mathbb{R}^{4})$ of 2-frames in $\mathbb{R}^{4}$. Notice also that the dimension $n=5$ and examples like the 5-sphere $S^{5}$, the product $S^{3}\times S^{2}$ and the space $M_{1, 1}$, are very important examples of 5-dimensional homogeneous or cohomogeneity-one Sasaki-Einstein manifolds. Finally, the sphere $S^{5}$ can be considered either as the irreducible symmetric space $SO(6)/SO(5)$, or as $SU(3)/SU(2)$ with a certain invariant metric. The isotropy representation of the second presentation decomposes into the sum of two irreducible summands. The proof of the fact that the canonical metric of $S^{5}=SU(3)/SU(2)$ is the unique $SU(3)$-invariant Einstein metric, was given by Jensen.</p> <p><strong>Difficulty of constructing general Einstein metrics:</strong> It is good to notice that Einstein metrics are relatively rare among Riemannian metrics. Although no topological obstructions are known for big dimensions, it is difficult to construct such metrics because the Einstein equation is a system of second order PDEs and there is no general method known. In recent decades, Einstein manifolds became the subject of many studies (both of mathematicians and physicists). Motivated from classical mechanics and the well known fact that symmetry considerations can simplify the study of geometric problems, the current strategy for constructing examples of Einstein metrics is to assume some symmetry condition for the metric tensor itself, that means that there is a Lie group which acts on the manifold in a smooth way by isometries. Roughly speaking, important progress has been done for homogeneous Riemannian $G$-manifolds and cohomogeneity-one Riemannian $G$-manifolds. The fist case arises from a transitive action of a Lie group $G$ on $(M, g)$, and the second one appears when the Lie group $G$ acts on $(M, g)$ with principal orbits of codimension one. Also, Einstein metrics with special holonomy is a famous topic (with a lot of interest in physics) and there are several results for this kind of Einstein metrics, e.g. Einstein-Sasakian manifolds, 6-dim nearly Kähler manifolds, 7-dim nearly parallel $G_2$-manifolds, etc, and all of them admit real Killing spinors. Also, Ricci-flatness is related with parallel spinors.</p> <p><strong>Finally:</strong> Possessing a first course in Riemmanian geometry and having some basic knowledge of representation theory... it is not difficult to prove for example that $SO(6)/SO(5)$ admits a unique $SO(6)$-invariant Einstein metric. But other problems related with Einstein metrics may need more expertise and studying. </p> Answer 2: <p>For what it is worth, there are many examples in <a href="https://arxiv.org/abs/1103.0817" rel="noreferrer">Dezhong Chen's paper.</a></p> Answer 3: <p>Simply connected 5-manifolds are classified by simple invariants, second homology and second Stiefel-Whitney class. So one can consider the existence problem, i.e. which 5-manifolds admit an Einstein metric. Many are know to have an Eintein metric. See the paper by János Kollár <a href="https://arxiv.org/abs/math/0408184" rel="nofollow noreferrer">"Einstein metrics on 5-dimensional Seifert bundles"</a>.</p> Answer 4: <p><strong>Answer for the compact homogeneous case:</strong> Due to Alekseevsky, Dotti and Ferraris (Homogeneous Ricci positive 5-manifolds, Pacific. J. Math, 175, 1-12, 1996) we know that in the non-symmetric case, the classification of 5-dimensional compact homogeneous Einstein spaces reduces to the classification of Einstein metrics on the space $M_{a, b}=SU(2)\times SU(2)/S^{1}$, with a specific embedding of $S^{1}$. Here, $a\geq b\geq 0$, $a&gt; 0$ are integers with ${\rm gcd}(a, b)&gt;1$. </p> <p>Wang and Ziller (Einstein metrics on principal torus bundles, J. Differ. Geom., 31, (1990) 215-248), proved that the spaces $M_{a, b}$ for $(a, b)\neq (1, 1)$ admit exactly one (up to isometry) homogeneous Einstein metric and the same result was obtained indepedently by Rodionov. Wang and Ziller also proved that $M_{1, 1}$ admits at least one invariant Einstein metric. Note that the isotropy representation of $M_{1, 1}$ decomposes into three irreducible submodules, two of them are isomorphic to each other, and thus a $SU(2)^{2}$-invariant Riemannian metric is depending on four positive real parameters. Now, $M_{1, 1}$ admits a second homogeneous Einstein metric, which is isometric to a product of symmetric metrics on $S^{3}\times S^{2}$. In fact, the space $M_{1, 1}$ can be viewed as the unit tangent bundle $T_1S^{3}=SO(4)/SO(2)$ of the 3-sphere $S^{3}$, and thus as the Stiefel manifold $V_{2}(\mathbb{R}^{4})$ of 2-frames in $\mathbb{R}^{4}$. Notice also that the dimension $n=5$ and examples like the 5-sphere $S^{5}$, the product $S^{3}\times S^{2}$ and the space $M_{1, 1}$, are very important examples of 5-dimensional homogeneous or cohomogeneity-one Sasaki-Einstein manifolds. Finally, the sphere $S^{5}$ can be considered either as the irreducible symmetric space $SO(6)/SO(5)$, or as $SU(3)/SU(2)$ with a certain invariant metric. The isotropy representation of the second presentation decomposes into the sum of two irreducible summands. The proof of the fact that the canonical metric of $S^{5}=SU(3)/SU(2)$ is the unique $SU(3)$-invariant Einstein metric, was given by Jensen.</p> <p><strong>Difficulty of constructing general Einstein metrics:</strong> It is good to notice that Einstein metrics are relatively rare among Riemannian metrics. Although no topological obstructions are known for big dimensions, it is difficult to construct such metrics because the Einstein equation is a system of second order PDEs and there is no general method known. In recent decades, Einstein manifolds became the subject of many studies (both of mathematicians and physicists). Motivated from classical mechanics and the well known fact that symmetry considerations can simplify the study of geometric problems, the current strategy for constructing examples of Einstein metrics is to assume some symmetry condition for the metric tensor itself, that means that there is a Lie group which acts on the manifold in a smooth way by isometries. Roughly speaking, important progress has been done for homogeneous Riemannian $G$-manifolds and cohomogeneity-one Riemannian $G$-manifolds. The fist case arises from a transitive action of a Lie group $G$ on $(M, g)$, and the second one appears when the Lie group $G$ acts on $(M, g)$ with principal orbits of codimension one. Also, Einstein metrics with special holonomy is a famous topic (with a lot of interest in physics) and there are several results for this kind of Einstein metrics, e.g. Einstein-Sasakian manifolds, 6-dim nearly Kähler manifolds, 7-dim nearly parallel $G_2$-manifolds, etc, and all of them admit real Killing spinors. Also, Ricci-flatness is related with parallel spinors.</p> <p><strong>Finally:</strong> Possessing a first course in Riemmanian geometry and having some basic knowledge of representation theory... it is not difficult to prove for example that $SO(6)/SO(5)$ admits a unique $SO(6)$-invariant Einstein metric. But other problems related with Einstein metrics may need more expertise and studying. </p> Answer 5: <p>For what it is worth, there are many examples in <a href="https://arxiv.org/abs/1103.0817" rel="noreferrer">Dezhong Chen's paper.</a></p> Answer 6: <p>Simply connected 5-manifolds are classified by simple invariants, second homology and second Stiefel-Whitney class. So one can consider the existence problem, i.e. which 5-manifolds admit an Einstein metric. Many are know to have an Eintein metric. See the paper by János Kollár <a href="https://arxiv.org/abs/math/0408184" rel="nofollow noreferrer">"Einstein metrics on 5-dimensional Seifert bundles"</a>.</p>
does this sum have a limit?
260,392
8
https://mathoverflow.net/questions/260392/does-this-sum-have-a-limit
<p>Define the sequence given by the finite sum <span class="math-container">$$a_n:=\sum_{k=2}^{n+1}\binom{2k}k\binom{n+1}k\frac{k-1}{2^k\binom{4n}k}.$$</span></p> <blockquote> <p><strong>Questions.</strong></p> <p>(1) Is <span class="math-container">$0&lt;a_n&lt;1$</span>?</p> <p>(2) Does the limit <span class="math-container">$\lim_{n\rightarrow\infty}a_n$</span> exist? If yes, what is its value?</p> </blockquote>
[ { "answer_id": 260399, "score": 5, "text": "<p>For each specific $k$ we have $$\\binom{2k}k\\binom{n+1}k\\frac{k-1}{2^k\\binom{4n}k}\\to \\binom{2k}k\\frac{k-1}{8^k},$$\nthat suggests that the sum tends to $$\\sum_{k=2}^\\infty \\binom{2k}k\\frac{k-1}{8^k}=1-\\frac{1}{\\sqrt 2},$$\nas $$\\sum_{k= 0}^\\infty \\binom{2k}k x^k =(1-4x)^{-1/2};\\,\\\\\\sum_{k= 0}^\\infty k\\binom{2k}k x^k =(x\\frac{d}{dx})(1-4x)^{-1/2}=2x(1-4x)^{-3/2};\\\\\n\\sum_{k= 0}^\\infty (k-1)\\binom{2k}k x^k =(6x-1)(1-4x)^{-3/2}=|_{x=1/8}-\\frac1{\\sqrt{2}}.$$</p>\n\n<p>In order to justify this limit changes (limit of a sum is a sum of limits) we majorate each term in our sum. Say, $$\\frac{\\binom{n+1}k}{\\binom{4n}k}=\\frac{(n+1)n(n-1)\\dots(n-k+2)}{4n(4n-1)(4n-2)\\dots (4n-k+1)}&lt;4^{-k}(1+\\frac{1}n)(1+\\frac{1}{4n-1})&lt;3\\cdot 4^{-k}.$$\nthis allows to estimate the total sum of summands with $k\\geqslant k_0$ by $3\\sum_{k\\geqslant k_0} \\binom{2k}k\\frac{k-1}{8^k}$, that tends to 0 for large $k_0$. Therefore the limit of your sum is indeed a sum of limits, moreover, the estimates are very specific and should prove that $0&lt;a_n&lt;1$ also after checking some small values of $n$.</p>\n" }, { "answer_id": 260395, "score": 3, "text": "<p>Not an answer. First, Mathematica says that \n$$a_n = \\frac{-4 n \\, _2F_1\\left(\\frac{1}{2},-n-1;-4 n;2\\right)+n \\,\n _2F_1\\left(\\frac{3}{2},-n;1-4 n;2\\right)+\\, _2F_1\\left(\\frac{3}{2},-n;1-4\n n;2\\right)}{4 n}.$$</p>\n\n<p>Secondly, for all reasonable values of $n$ it seems that $a_n &lt; -0.7,$ and, indeed, I conjecture that \n$$\\lim_{n\\rightarrow \\infty} a_n = -\\frac{1}{\\sqrt{2}}.$$</p>\n\n<p>Now, this is a response to the original form of this question, with the sum going from $0.$ The first two terms contribute $-1$ to the sum, so for the current version, the conjectured limit is \n$$1-\\frac{1}{\\sqrt{2}}.$$</p>\n" }, { "answer_id": 260399, "score": 5, "text": "<p>For each specific $k$ we have $$\\binom{2k}k\\binom{n+1}k\\frac{k-1}{2^k\\binom{4n}k}\\to \\binom{2k}k\\frac{k-1}{8^k},$$\nthat suggests that the sum tends to $$\\sum_{k=2}^\\infty \\binom{2k}k\\frac{k-1}{8^k}=1-\\frac{1}{\\sqrt 2},$$\nas $$\\sum_{k= 0}^\\infty \\binom{2k}k x^k =(1-4x)^{-1/2};\\,\\\\\\sum_{k= 0}^\\infty k\\binom{2k}k x^k =(x\\frac{d}{dx})(1-4x)^{-1/2}=2x(1-4x)^{-3/2};\\\\\n\\sum_{k= 0}^\\infty (k-1)\\binom{2k}k x^k =(6x-1)(1-4x)^{-3/2}=|_{x=1/8}-\\frac1{\\sqrt{2}}.$$</p>\n\n<p>In order to justify this limit changes (limit of a sum is a sum of limits) we majorate each term in our sum. Say, $$\\frac{\\binom{n+1}k}{\\binom{4n}k}=\\frac{(n+1)n(n-1)\\dots(n-k+2)}{4n(4n-1)(4n-2)\\dots (4n-k+1)}&lt;4^{-k}(1+\\frac{1}n)(1+\\frac{1}{4n-1})&lt;3\\cdot 4^{-k}.$$\nthis allows to estimate the total sum of summands with $k\\geqslant k_0$ by $3\\sum_{k\\geqslant k_0} \\binom{2k}k\\frac{k-1}{8^k}$, that tends to 0 for large $k_0$. Therefore the limit of your sum is indeed a sum of limits, moreover, the estimates are very specific and should prove that $0&lt;a_n&lt;1$ also after checking some small values of $n$.</p>\n" }, { "answer_id": 260395, "score": 3, "text": "<p>Not an answer. First, Mathematica says that \n$$a_n = \\frac{-4 n \\, _2F_1\\left(\\frac{1}{2},-n-1;-4 n;2\\right)+n \\,\n _2F_1\\left(\\frac{3}{2},-n;1-4 n;2\\right)+\\, _2F_1\\left(\\frac{3}{2},-n;1-4\n n;2\\right)}{4 n}.$$</p>\n\n<p>Secondly, for all reasonable values of $n$ it seems that $a_n &lt; -0.7,$ and, indeed, I conjecture that \n$$\\lim_{n\\rightarrow \\infty} a_n = -\\frac{1}{\\sqrt{2}}.$$</p>\n\n<p>Now, this is a response to the original form of this question, with the sum going from $0.$ The first two terms contribute $-1$ to the sum, so for the current version, the conjectured limit is \n$$1-\\frac{1}{\\sqrt{2}}.$$</p>\n" } ]
Question: does this sum have a limit? Answer 1: <p>For each specific $k$ we have $$\binom{2k}k\binom{n+1}k\frac{k-1}{2^k\binom{4n}k}\to \binom{2k}k\frac{k-1}{8^k},$$ that suggests that the sum tends to $$\sum_{k=2}^\infty \binom{2k}k\frac{k-1}{8^k}=1-\frac{1}{\sqrt 2},$$ as $$\sum_{k= 0}^\infty \binom{2k}k x^k =(1-4x)^{-1/2};\,\\\sum_{k= 0}^\infty k\binom{2k}k x^k =(x\frac{d}{dx})(1-4x)^{-1/2}=2x(1-4x)^{-3/2};\\ \sum_{k= 0}^\infty (k-1)\binom{2k}k x^k =(6x-1)(1-4x)^{-3/2}=|_{x=1/8}-\frac1{\sqrt{2}}.$$</p> <p>In order to justify this limit changes (limit of a sum is a sum of limits) we majorate each term in our sum. Say, $$\frac{\binom{n+1}k}{\binom{4n}k}=\frac{(n+1)n(n-1)\dots(n-k+2)}{4n(4n-1)(4n-2)\dots (4n-k+1)}&lt;4^{-k}(1+\frac{1}n)(1+\frac{1}{4n-1})&lt;3\cdot 4^{-k}.$$ this allows to estimate the total sum of summands with $k\geqslant k_0$ by $3\sum_{k\geqslant k_0} \binom{2k}k\frac{k-1}{8^k}$, that tends to 0 for large $k_0$. Therefore the limit of your sum is indeed a sum of limits, moreover, the estimates are very specific and should prove that $0&lt;a_n&lt;1$ also after checking some small values of $n$.</p> Answer 2: <p>Not an answer. First, Mathematica says that $$a_n = \frac{-4 n \, _2F_1\left(\frac{1}{2},-n-1;-4 n;2\right)+n \, _2F_1\left(\frac{3}{2},-n;1-4 n;2\right)+\, _2F_1\left(\frac{3}{2},-n;1-4 n;2\right)}{4 n}.$$</p> <p>Secondly, for all reasonable values of $n$ it seems that $a_n &lt; -0.7,$ and, indeed, I conjecture that $$\lim_{n\rightarrow \infty} a_n = -\frac{1}{\sqrt{2}}.$$</p> <p>Now, this is a response to the original form of this question, with the sum going from $0.$ The first two terms contribute $-1$ to the sum, so for the current version, the conjectured limit is $$1-\frac{1}{\sqrt{2}}.$$</p> Answer 3: <p>For each specific $k$ we have $$\binom{2k}k\binom{n+1}k\frac{k-1}{2^k\binom{4n}k}\to \binom{2k}k\frac{k-1}{8^k},$$ that suggests that the sum tends to $$\sum_{k=2}^\infty \binom{2k}k\frac{k-1}{8^k}=1-\frac{1}{\sqrt 2},$$ as $$\sum_{k= 0}^\infty \binom{2k}k x^k =(1-4x)^{-1/2};\,\\\sum_{k= 0}^\infty k\binom{2k}k x^k =(x\frac{d}{dx})(1-4x)^{-1/2}=2x(1-4x)^{-3/2};\\ \sum_{k= 0}^\infty (k-1)\binom{2k}k x^k =(6x-1)(1-4x)^{-3/2}=|_{x=1/8}-\frac1{\sqrt{2}}.$$</p> <p>In order to justify this limit changes (limit of a sum is a sum of limits) we majorate each term in our sum. Say, $$\frac{\binom{n+1}k}{\binom{4n}k}=\frac{(n+1)n(n-1)\dots(n-k+2)}{4n(4n-1)(4n-2)\dots (4n-k+1)}&lt;4^{-k}(1+\frac{1}n)(1+\frac{1}{4n-1})&lt;3\cdot 4^{-k}.$$ this allows to estimate the total sum of summands with $k\geqslant k_0$ by $3\sum_{k\geqslant k_0} \binom{2k}k\frac{k-1}{8^k}$, that tends to 0 for large $k_0$. Therefore the limit of your sum is indeed a sum of limits, moreover, the estimates are very specific and should prove that $0&lt;a_n&lt;1$ also after checking some small values of $n$.</p> Answer 4: <p>Not an answer. First, Mathematica says that $$a_n = \frac{-4 n \, _2F_1\left(\frac{1}{2},-n-1;-4 n;2\right)+n \, _2F_1\left(\frac{3}{2},-n;1-4 n;2\right)+\, _2F_1\left(\frac{3}{2},-n;1-4 n;2\right)}{4 n}.$$</p> <p>Secondly, for all reasonable values of $n$ it seems that $a_n &lt; -0.7,$ and, indeed, I conjecture that $$\lim_{n\rightarrow \infty} a_n = -\frac{1}{\sqrt{2}}.$$</p> <p>Now, this is a response to the original form of this question, with the sum going from $0.$ The first two terms contribute $-1$ to the sum, so for the current version, the conjectured limit is $$1-\frac{1}{\sqrt{2}}.$$</p>
Sum of square roots of binomial coefficients
258,711
7
https://mathoverflow.net/questions/258711/sum-of-square-roots-of-binomial-coefficients
<p><strong>Question:</strong> let \begin{equation} c_n=\sum_{i=0}^n \sqrt{\binom{n}{i}}. \end{equation} How does $c_n$ grow with $n$?</p> <p>My conjecture is that $c_n=\Theta(2^{0.5n}n^{0.25})$.This is because \begin{equation} 0.5^{0.5n} c_n = \sum_{i=0}^n \sqrt{\binom{n}{i}0.5^n}=\sum_{i=0}^n\sqrt{Pr(X=i)}, \end{equation} for $X\sim Bin(n,0.5)$. The binomial distribution $Bin(n,0.5)$ is approximately the normal distribution $N(0.5n, 0.25n)$. Also, if $Y\sim N(0.5n, 0.25n)$, it is not hard to see that \begin{equation} \int \sqrt{f(y)} dy =\Theta(n^{0.25}). \end{equation} Therefore I believe that it is also true that $0.5^{0.5n} c_n=\Theta(n^{0.25})$. I played with matlab to get some evidence and found that \begin{equation} \frac{0.5^{0.5n} c_n}{n^{0.25}} \approx \frac{\pi}{2}, \end{equation} for $n=100,1000,10000,100000,1000000$. So the conjecture should be true. So anyone can prove it?</p> <p>(It turns out that $\frac{0.5^{0.5n} c_n}{n^{0.25}} \approx (2\pi)^{0.25}$ instead of $\frac{\pi}{2}$).</p>
[ { "answer_id": 258719, "score": 9, "text": "<p>For smallish $k$, we have\n$$ \\binom{n}{n/2+k} \\approx \\binom{n}{n/2} \\exp(-2k^2/n). $$\nSo\n$$\\sum\\sqrt{\\binom{n}{n/2+k}}\n \\approx \\sqrt{\\binom{n}{n/2}} \\int_{-\\infty}^\\infty e^{-k^2/n}\\,dk\n \\approx 2^{n/2} (2\\pi n)^{1/4}.$$\nOf course I did some hand-waving here, but all of this is rigorously justifiable. Use the Euler-Maclaurin theorm to justify replacing the sum by an integral.</p>\n" }, { "answer_id": 258713, "score": 4, "text": "<p>Define the complex-variable function\n$$g(z)=\\frac{\\sin\\pi z}{\\pi z}\\prod_{k=1}^n\\frac1{1-\\frac{z}k}.$$\nNote $f(k)=\\binom{n}k$ for $k\\in\\mathbb{Z}^{\\geq0}$. Moreover, $\\sqrt{g(z)}$ is analytic in the region $-1&lt;\\Re(z)&lt;n+1$ since $g(z)$ is zero-free there. By the <a href=\"https://en.wikipedia.org/wiki/Residue_theorem\" rel=\"nofollow noreferrer\">Residue Theorem</a>,\n$$\\sum_{k=0}^n\\sqrt{\\binom{n}k}\n=\\frac1{2\\pi i}\\int_C\\sqrt{g(z)}\\,\\frac{\\pi\\cos\\pi z}{\\sin\\pi z}\\,dz$$\nwhere $C$ is a closed (positive) contour inside $-1&lt;\\Re(z)&lt;n+1$ going around once about each integer $0\\leq k\\leq n$. Choose $C_R$ as the rectangular along the lines $\\Re(z)=-\\frac12$, $\\Re(z)=n+\\frac12$ and $\\Im(z)=\\pm R$. Thus\n$$\\int_{C_R}\\sqrt{g(z)}\\,\\frac{\\pi\\cos\\pi z}{\\sin\\pi z}\\,dz\n=\\int_{C_R}\\prod_{k=1}^n\\sqrt{\\frac1{1-\\frac{z}k}}\\,\\frac{\\sqrt{\\pi}\\cos\\pi z}\n{\\sqrt{z\\sin\\pi z}}\\,dz.$$\nSo, letting $R\\rightarrow\\infty$, we arrive at\n$$\\sum_{k=0}^n\\sqrt{\\binom{n}k}\n=\\frac1{2\\pi i}\\left[\\int_{-\\frac12+i\\infty}^{-\\frac12-i\\infty}\n+\\int_{n+\\frac12-i\\infty}^{n+\\frac12+i\\infty}\n\\sqrt{g(z)}\\,\\frac{\\pi\\cos\\pi z}{\\sin\\pi z}\\,dz\\right].$$\nThe integrand is invariant under $z\\rightarrow n-z$, hence it suffices to compute (actually estimate) one of the integrals.</p>\n\n<p>This is where I left off. Perhaps someone can complete the argument.</p>\n" }, { "answer_id": 258719, "score": 9, "text": "<p>For smallish $k$, we have\n$$ \\binom{n}{n/2+k} \\approx \\binom{n}{n/2} \\exp(-2k^2/n). $$\nSo\n$$\\sum\\sqrt{\\binom{n}{n/2+k}}\n \\approx \\sqrt{\\binom{n}{n/2}} \\int_{-\\infty}^\\infty e^{-k^2/n}\\,dk\n \\approx 2^{n/2} (2\\pi n)^{1/4}.$$\nOf course I did some hand-waving here, but all of this is rigorously justifiable. Use the Euler-Maclaurin theorm to justify replacing the sum by an integral.</p>\n" }, { "answer_id": 258713, "score": 4, "text": "<p>Define the complex-variable function\n$$g(z)=\\frac{\\sin\\pi z}{\\pi z}\\prod_{k=1}^n\\frac1{1-\\frac{z}k}.$$\nNote $f(k)=\\binom{n}k$ for $k\\in\\mathbb{Z}^{\\geq0}$. Moreover, $\\sqrt{g(z)}$ is analytic in the region $-1&lt;\\Re(z)&lt;n+1$ since $g(z)$ is zero-free there. By the <a href=\"https://en.wikipedia.org/wiki/Residue_theorem\" rel=\"nofollow noreferrer\">Residue Theorem</a>,\n$$\\sum_{k=0}^n\\sqrt{\\binom{n}k}\n=\\frac1{2\\pi i}\\int_C\\sqrt{g(z)}\\,\\frac{\\pi\\cos\\pi z}{\\sin\\pi z}\\,dz$$\nwhere $C$ is a closed (positive) contour inside $-1&lt;\\Re(z)&lt;n+1$ going around once about each integer $0\\leq k\\leq n$. Choose $C_R$ as the rectangular along the lines $\\Re(z)=-\\frac12$, $\\Re(z)=n+\\frac12$ and $\\Im(z)=\\pm R$. Thus\n$$\\int_{C_R}\\sqrt{g(z)}\\,\\frac{\\pi\\cos\\pi z}{\\sin\\pi z}\\,dz\n=\\int_{C_R}\\prod_{k=1}^n\\sqrt{\\frac1{1-\\frac{z}k}}\\,\\frac{\\sqrt{\\pi}\\cos\\pi z}\n{\\sqrt{z\\sin\\pi z}}\\,dz.$$\nSo, letting $R\\rightarrow\\infty$, we arrive at\n$$\\sum_{k=0}^n\\sqrt{\\binom{n}k}\n=\\frac1{2\\pi i}\\left[\\int_{-\\frac12+i\\infty}^{-\\frac12-i\\infty}\n+\\int_{n+\\frac12-i\\infty}^{n+\\frac12+i\\infty}\n\\sqrt{g(z)}\\,\\frac{\\pi\\cos\\pi z}{\\sin\\pi z}\\,dz\\right].$$\nThe integrand is invariant under $z\\rightarrow n-z$, hence it suffices to compute (actually estimate) one of the integrals.</p>\n\n<p>This is where I left off. Perhaps someone can complete the argument.</p>\n" } ]
Question: Sum of square roots of binomial coefficients Answer 1: <p>For smallish $k$, we have $$ \binom{n}{n/2+k} \approx \binom{n}{n/2} \exp(-2k^2/n). $$ So $$\sum\sqrt{\binom{n}{n/2+k}} \approx \sqrt{\binom{n}{n/2}} \int_{-\infty}^\infty e^{-k^2/n}\,dk \approx 2^{n/2} (2\pi n)^{1/4}.$$ Of course I did some hand-waving here, but all of this is rigorously justifiable. Use the Euler-Maclaurin theorm to justify replacing the sum by an integral.</p> Answer 2: <p>Define the complex-variable function $$g(z)=\frac{\sin\pi z}{\pi z}\prod_{k=1}^n\frac1{1-\frac{z}k}.$$ Note $f(k)=\binom{n}k$ for $k\in\mathbb{Z}^{\geq0}$. Moreover, $\sqrt{g(z)}$ is analytic in the region $-1&lt;\Re(z)&lt;n+1$ since $g(z)$ is zero-free there. By the <a href="https://en.wikipedia.org/wiki/Residue_theorem" rel="nofollow noreferrer">Residue Theorem</a>, $$\sum_{k=0}^n\sqrt{\binom{n}k} =\frac1{2\pi i}\int_C\sqrt{g(z)}\,\frac{\pi\cos\pi z}{\sin\pi z}\,dz$$ where $C$ is a closed (positive) contour inside $-1&lt;\Re(z)&lt;n+1$ going around once about each integer $0\leq k\leq n$. Choose $C_R$ as the rectangular along the lines $\Re(z)=-\frac12$, $\Re(z)=n+\frac12$ and $\Im(z)=\pm R$. Thus $$\int_{C_R}\sqrt{g(z)}\,\frac{\pi\cos\pi z}{\sin\pi z}\,dz =\int_{C_R}\prod_{k=1}^n\sqrt{\frac1{1-\frac{z}k}}\,\frac{\sqrt{\pi}\cos\pi z} {\sqrt{z\sin\pi z}}\,dz.$$ So, letting $R\rightarrow\infty$, we arrive at $$\sum_{k=0}^n\sqrt{\binom{n}k} =\frac1{2\pi i}\left[\int_{-\frac12+i\infty}^{-\frac12-i\infty} +\int_{n+\frac12-i\infty}^{n+\frac12+i\infty} \sqrt{g(z)}\,\frac{\pi\cos\pi z}{\sin\pi z}\,dz\right].$$ The integrand is invariant under $z\rightarrow n-z$, hence it suffices to compute (actually estimate) one of the integrals.</p> <p>This is where I left off. Perhaps someone can complete the argument.</p> Answer 3: <p>For smallish $k$, we have $$ \binom{n}{n/2+k} \approx \binom{n}{n/2} \exp(-2k^2/n). $$ So $$\sum\sqrt{\binom{n}{n/2+k}} \approx \sqrt{\binom{n}{n/2}} \int_{-\infty}^\infty e^{-k^2/n}\,dk \approx 2^{n/2} (2\pi n)^{1/4}.$$ Of course I did some hand-waving here, but all of this is rigorously justifiable. Use the Euler-Maclaurin theorm to justify replacing the sum by an integral.</p> Answer 4: <p>Define the complex-variable function $$g(z)=\frac{\sin\pi z}{\pi z}\prod_{k=1}^n\frac1{1-\frac{z}k}.$$ Note $f(k)=\binom{n}k$ for $k\in\mathbb{Z}^{\geq0}$. Moreover, $\sqrt{g(z)}$ is analytic in the region $-1&lt;\Re(z)&lt;n+1$ since $g(z)$ is zero-free there. By the <a href="https://en.wikipedia.org/wiki/Residue_theorem" rel="nofollow noreferrer">Residue Theorem</a>, $$\sum_{k=0}^n\sqrt{\binom{n}k} =\frac1{2\pi i}\int_C\sqrt{g(z)}\,\frac{\pi\cos\pi z}{\sin\pi z}\,dz$$ where $C$ is a closed (positive) contour inside $-1&lt;\Re(z)&lt;n+1$ going around once about each integer $0\leq k\leq n$. Choose $C_R$ as the rectangular along the lines $\Re(z)=-\frac12$, $\Re(z)=n+\frac12$ and $\Im(z)=\pm R$. Thus $$\int_{C_R}\sqrt{g(z)}\,\frac{\pi\cos\pi z}{\sin\pi z}\,dz =\int_{C_R}\prod_{k=1}^n\sqrt{\frac1{1-\frac{z}k}}\,\frac{\sqrt{\pi}\cos\pi z} {\sqrt{z\sin\pi z}}\,dz.$$ So, letting $R\rightarrow\infty$, we arrive at $$\sum_{k=0}^n\sqrt{\binom{n}k} =\frac1{2\pi i}\left[\int_{-\frac12+i\infty}^{-\frac12-i\infty} +\int_{n+\frac12-i\infty}^{n+\frac12+i\infty} \sqrt{g(z)}\,\frac{\pi\cos\pi z}{\sin\pi z}\,dz\right].$$ The integrand is invariant under $z\rightarrow n-z$, hence it suffices to compute (actually estimate) one of the integrals.</p> <p>This is where I left off. Perhaps someone can complete the argument.</p>
Finite surjective morphism of normal varieties and Galois coverings
258,747
7
https://mathoverflow.net/questions/258747/finite-surjective-morphism-of-normal-varieties-and-galois-coverings
<p>Let $f:X\to Y$ be a finite surjective morphism of normal varieties. We say that $f$ is a Galois covering if the extension $k(X)/k(Y)$ is Galois.</p> <p>I have the following questions:</p> <p>1.- Why we need to assume normality of the varieties for the definition of a Galois covering?</p> <p>2.- Is it true that a finite surjective morphism between normal varieties is always flat? </p> <p>3.- Is it true that $f$ is a Galois covering if and only if the group of automorphism $Aut(X/Y)$ acts transitively on all fibers of $f$?</p>
[ { "answer_id": 258749, "score": 13, "text": "<p>The simplest counterexample to question 2 is the map \n$$\nX := \\mathbb{A}^2 \\to \\mathbb{A}^2/\\{ \\pm 1 \\} =: Y.\n$$\nHere $Y$ is a quadratic cone in $\\mathbb{A}^3$. Both $X$ and $Y$ are normal, and the map is a Galois covering, but $X$ is not flat over $Y$.</p>\n" }, { "answer_id": 258748, "score": 10, "text": "<p>(I assume by a covering, you mean a finite surjective <em>étale</em> morphism.)</p>\n\n<ol>\n<li><p>This is needed as in your definition of Galois covering you only look at the extension of the function fields. Using the more general definition from [SGA1, Exp. V, Remarque 5.11], this works for any connected scheme. For this, see also [Stacks Project], <a href=\"http://stacks.math.columbia.edu/download/pione.pdf\" rel=\"nofollow noreferrer\">http://stacks.math.columbia.edu/download/pione.pdf</a>, p. 6, after Lemma 3.7: $X/Y$ is Galois (this works for any Galois category with fibre functor $\\mathrm{Fib}_\\bar{y}$) iff $X$ is connected and $|\\mathrm{Aut}(X/Y)| = \\mathrm{Fib}_\\bar{y}X$ iff $X$ is connected and $\\mathrm{Aut}(X/Y)$ acts transitively on $\\mathrm{Fib}_\\bar{y}X$. The étale fundamental group is defined in Definition 6.1 there. In section 10, the fundamental group of <em>normal</em> schemes is considered. See also <a href=\"http://websites.math.leidenuniv.nl/algebra/GSchemes.pdf\" rel=\"nofollow noreferrer\">http://websites.math.leidenuniv.nl/algebra/GSchemes.pdf</a>, p. 41, 3.14 Galois objects. For the question what happens in the non-normal case, see <a href=\"https://mathoverflow.net/questions/129703/the-%C3%A9tale-fundamental-group-in-the-non-normal-case\">The &#233;tale fundamental group in the non-normal case</a>.</p></li>\n<li><p>If $f: X \\to Y$ is a finite surjective morphism of <em>regular</em> schemes (or more generally, if $Y$ is regular and $X$ is Cohen-Macaulay), it is flat, see [Liu], Remark 4.3.11, or [EGAIV${}_3$], p. 230, Proposition (15.4.2).</p></li>\n<li><p>I think this is true, but I do not have a reference. I would show that $\\mathrm{Aut}(X/Y)$ acting transitively on the fibres is equivalent to the usual definition of a Galois cover from [SGA1], and then that this is equivalent to the generic fibre being a Galois cover. I think this follows from the fact that one has a surjection $G_{k(Y)} \\twoheadrightarrow \\pi_1^{\\mathrm{\\acute{e}t}}(Y,\\bar{y})$ and under surjective group homomorphisms, normal subgroups correspond to normal subgroups.</p></li>\n</ol>\n" }, { "answer_id": 258764, "score": 9, "text": "<p>I'm just going to make a quick additional comment on #2 (just to add on to Timo Keller's explanations and Sasha's counter example).</p>\n\n<p><strong>Theorem (see Matsumura, page 179)</strong> <em>If $R \\subseteq S$ is a finite extension of Noetherian rings with $R$ regular and $S$ Cohen-Macaulay, then $S$ is a flat $R$ module.</em></p>\n\n<p>So you just need $Y$ to be regular and $X$ to be Cohen-Macaulay (Matsumura actually has a more general statement which also works in the non-finite case).</p>\n\n<p>On the other hand, if $Y$ is <em>not</em> regular then $X/Y$ is very rarely flat. Indeed, since the map is finite, flat is the same as $f_* O_Y$ being a locally free $O_X$-module and we can ask how many (local) $O_X$-summands $f_* O_Y$ has as an $O_X$-module.</p>\n\n<p><strong>Fact</strong> If $R \\subseteq S$ is a finite extension of normal local rings with the same residue field which is etale in codimension 1 (ie, the sort of group quotient that Sasha wrote above), then $S$ has at most one $R$-summand as an $R$-module (in particular, it is nowhere near flat). See the question I asked earlier: <a href=\"https://mathoverflow.net/questions/236048/number-of-free-summands-of-finite-local-extensions\">Number of free summands of finite local extensions</a></p>\n" }, { "answer_id": 258749, "score": 13, "text": "<p>The simplest counterexample to question 2 is the map \n$$\nX := \\mathbb{A}^2 \\to \\mathbb{A}^2/\\{ \\pm 1 \\} =: Y.\n$$\nHere $Y$ is a quadratic cone in $\\mathbb{A}^3$. Both $X$ and $Y$ are normal, and the map is a Galois covering, but $X$ is not flat over $Y$.</p>\n" }, { "answer_id": 258748, "score": 10, "text": "<p>(I assume by a covering, you mean a finite surjective <em>étale</em> morphism.)</p>\n\n<ol>\n<li><p>This is needed as in your definition of Galois covering you only look at the extension of the function fields. Using the more general definition from [SGA1, Exp. V, Remarque 5.11], this works for any connected scheme. For this, see also [Stacks Project], <a href=\"http://stacks.math.columbia.edu/download/pione.pdf\" rel=\"nofollow noreferrer\">http://stacks.math.columbia.edu/download/pione.pdf</a>, p. 6, after Lemma 3.7: $X/Y$ is Galois (this works for any Galois category with fibre functor $\\mathrm{Fib}_\\bar{y}$) iff $X$ is connected and $|\\mathrm{Aut}(X/Y)| = \\mathrm{Fib}_\\bar{y}X$ iff $X$ is connected and $\\mathrm{Aut}(X/Y)$ acts transitively on $\\mathrm{Fib}_\\bar{y}X$. The étale fundamental group is defined in Definition 6.1 there. In section 10, the fundamental group of <em>normal</em> schemes is considered. See also <a href=\"http://websites.math.leidenuniv.nl/algebra/GSchemes.pdf\" rel=\"nofollow noreferrer\">http://websites.math.leidenuniv.nl/algebra/GSchemes.pdf</a>, p. 41, 3.14 Galois objects. For the question what happens in the non-normal case, see <a href=\"https://mathoverflow.net/questions/129703/the-%C3%A9tale-fundamental-group-in-the-non-normal-case\">The &#233;tale fundamental group in the non-normal case</a>.</p></li>\n<li><p>If $f: X \\to Y$ is a finite surjective morphism of <em>regular</em> schemes (or more generally, if $Y$ is regular and $X$ is Cohen-Macaulay), it is flat, see [Liu], Remark 4.3.11, or [EGAIV${}_3$], p. 230, Proposition (15.4.2).</p></li>\n<li><p>I think this is true, but I do not have a reference. I would show that $\\mathrm{Aut}(X/Y)$ acting transitively on the fibres is equivalent to the usual definition of a Galois cover from [SGA1], and then that this is equivalent to the generic fibre being a Galois cover. I think this follows from the fact that one has a surjection $G_{k(Y)} \\twoheadrightarrow \\pi_1^{\\mathrm{\\acute{e}t}}(Y,\\bar{y})$ and under surjective group homomorphisms, normal subgroups correspond to normal subgroups.</p></li>\n</ol>\n" }, { "answer_id": 258764, "score": 9, "text": "<p>I'm just going to make a quick additional comment on #2 (just to add on to Timo Keller's explanations and Sasha's counter example).</p>\n\n<p><strong>Theorem (see Matsumura, page 179)</strong> <em>If $R \\subseteq S$ is a finite extension of Noetherian rings with $R$ regular and $S$ Cohen-Macaulay, then $S$ is a flat $R$ module.</em></p>\n\n<p>So you just need $Y$ to be regular and $X$ to be Cohen-Macaulay (Matsumura actually has a more general statement which also works in the non-finite case).</p>\n\n<p>On the other hand, if $Y$ is <em>not</em> regular then $X/Y$ is very rarely flat. Indeed, since the map is finite, flat is the same as $f_* O_Y$ being a locally free $O_X$-module and we can ask how many (local) $O_X$-summands $f_* O_Y$ has as an $O_X$-module.</p>\n\n<p><strong>Fact</strong> If $R \\subseteq S$ is a finite extension of normal local rings with the same residue field which is etale in codimension 1 (ie, the sort of group quotient that Sasha wrote above), then $S$ has at most one $R$-summand as an $R$-module (in particular, it is nowhere near flat). See the question I asked earlier: <a href=\"https://mathoverflow.net/questions/236048/number-of-free-summands-of-finite-local-extensions\">Number of free summands of finite local extensions</a></p>\n" } ]
Question: Finite surjective morphism of normal varieties and Galois coverings Answer 1: <p>The simplest counterexample to question 2 is the map $$ X := \mathbb{A}^2 \to \mathbb{A}^2/\{ \pm 1 \} =: Y. $$ Here $Y$ is a quadratic cone in $\mathbb{A}^3$. Both $X$ and $Y$ are normal, and the map is a Galois covering, but $X$ is not flat over $Y$.</p> Answer 2: <p>(I assume by a covering, you mean a finite surjective <em>étale</em> morphism.)</p> <ol> <li><p>This is needed as in your definition of Galois covering you only look at the extension of the function fields. Using the more general definition from [SGA1, Exp. V, Remarque 5.11], this works for any connected scheme. For this, see also [Stacks Project], <a href="http://stacks.math.columbia.edu/download/pione.pdf" rel="nofollow noreferrer">http://stacks.math.columbia.edu/download/pione.pdf</a>, p. 6, after Lemma 3.7: $X/Y$ is Galois (this works for any Galois category with fibre functor $\mathrm{Fib}_\bar{y}$) iff $X$ is connected and $|\mathrm{Aut}(X/Y)| = \mathrm{Fib}_\bar{y}X$ iff $X$ is connected and $\mathrm{Aut}(X/Y)$ acts transitively on $\mathrm{Fib}_\bar{y}X$. The étale fundamental group is defined in Definition 6.1 there. In section 10, the fundamental group of <em>normal</em> schemes is considered. See also <a href="http://websites.math.leidenuniv.nl/algebra/GSchemes.pdf" rel="nofollow noreferrer">http://websites.math.leidenuniv.nl/algebra/GSchemes.pdf</a>, p. 41, 3.14 Galois objects. For the question what happens in the non-normal case, see <a href="https://mathoverflow.net/questions/129703/the-%C3%A9tale-fundamental-group-in-the-non-normal-case">The &#233;tale fundamental group in the non-normal case</a>.</p></li> <li><p>If $f: X \to Y$ is a finite surjective morphism of <em>regular</em> schemes (or more generally, if $Y$ is regular and $X$ is Cohen-Macaulay), it is flat, see [Liu], Remark 4.3.11, or [EGAIV${}_3$], p. 230, Proposition (15.4.2).</p></li> <li><p>I think this is true, but I do not have a reference. I would show that $\mathrm{Aut}(X/Y)$ acting transitively on the fibres is equivalent to the usual definition of a Galois cover from [SGA1], and then that this is equivalent to the generic fibre being a Galois cover. I think this follows from the fact that one has a surjection $G_{k(Y)} \twoheadrightarrow \pi_1^{\mathrm{\acute{e}t}}(Y,\bar{y})$ and under surjective group homomorphisms, normal subgroups correspond to normal subgroups.</p></li> </ol> Answer 3: <p>I'm just going to make a quick additional comment on #2 (just to add on to Timo Keller's explanations and Sasha's counter example).</p> <p><strong>Theorem (see Matsumura, page 179)</strong> <em>If $R \subseteq S$ is a finite extension of Noetherian rings with $R$ regular and $S$ Cohen-Macaulay, then $S$ is a flat $R$ module.</em></p> <p>So you just need $Y$ to be regular and $X$ to be Cohen-Macaulay (Matsumura actually has a more general statement which also works in the non-finite case).</p> <p>On the other hand, if $Y$ is <em>not</em> regular then $X/Y$ is very rarely flat. Indeed, since the map is finite, flat is the same as $f_* O_Y$ being a locally free $O_X$-module and we can ask how many (local) $O_X$-summands $f_* O_Y$ has as an $O_X$-module.</p> <p><strong>Fact</strong> If $R \subseteq S$ is a finite extension of normal local rings with the same residue field which is etale in codimension 1 (ie, the sort of group quotient that Sasha wrote above), then $S$ has at most one $R$-summand as an $R$-module (in particular, it is nowhere near flat). See the question I asked earlier: <a href="https://mathoverflow.net/questions/236048/number-of-free-summands-of-finite-local-extensions">Number of free summands of finite local extensions</a></p> Answer 4: <p>The simplest counterexample to question 2 is the map $$ X := \mathbb{A}^2 \to \mathbb{A}^2/\{ \pm 1 \} =: Y. $$ Here $Y$ is a quadratic cone in $\mathbb{A}^3$. Both $X$ and $Y$ are normal, and the map is a Galois covering, but $X$ is not flat over $Y$.</p> Answer 5: <p>(I assume by a covering, you mean a finite surjective <em>étale</em> morphism.)</p> <ol> <li><p>This is needed as in your definition of Galois covering you only look at the extension of the function fields. Using the more general definition from [SGA1, Exp. V, Remarque 5.11], this works for any connected scheme. For this, see also [Stacks Project], <a href="http://stacks.math.columbia.edu/download/pione.pdf" rel="nofollow noreferrer">http://stacks.math.columbia.edu/download/pione.pdf</a>, p. 6, after Lemma 3.7: $X/Y$ is Galois (this works for any Galois category with fibre functor $\mathrm{Fib}_\bar{y}$) iff $X$ is connected and $|\mathrm{Aut}(X/Y)| = \mathrm{Fib}_\bar{y}X$ iff $X$ is connected and $\mathrm{Aut}(X/Y)$ acts transitively on $\mathrm{Fib}_\bar{y}X$. The étale fundamental group is defined in Definition 6.1 there. In section 10, the fundamental group of <em>normal</em> schemes is considered. See also <a href="http://websites.math.leidenuniv.nl/algebra/GSchemes.pdf" rel="nofollow noreferrer">http://websites.math.leidenuniv.nl/algebra/GSchemes.pdf</a>, p. 41, 3.14 Galois objects. For the question what happens in the non-normal case, see <a href="https://mathoverflow.net/questions/129703/the-%C3%A9tale-fundamental-group-in-the-non-normal-case">The &#233;tale fundamental group in the non-normal case</a>.</p></li> <li><p>If $f: X \to Y$ is a finite surjective morphism of <em>regular</em> schemes (or more generally, if $Y$ is regular and $X$ is Cohen-Macaulay), it is flat, see [Liu], Remark 4.3.11, or [EGAIV${}_3$], p. 230, Proposition (15.4.2).</p></li> <li><p>I think this is true, but I do not have a reference. I would show that $\mathrm{Aut}(X/Y)$ acting transitively on the fibres is equivalent to the usual definition of a Galois cover from [SGA1], and then that this is equivalent to the generic fibre being a Galois cover. I think this follows from the fact that one has a surjection $G_{k(Y)} \twoheadrightarrow \pi_1^{\mathrm{\acute{e}t}}(Y,\bar{y})$ and under surjective group homomorphisms, normal subgroups correspond to normal subgroups.</p></li> </ol> Answer 6: <p>I'm just going to make a quick additional comment on #2 (just to add on to Timo Keller's explanations and Sasha's counter example).</p> <p><strong>Theorem (see Matsumura, page 179)</strong> <em>If $R \subseteq S$ is a finite extension of Noetherian rings with $R$ regular and $S$ Cohen-Macaulay, then $S$ is a flat $R$ module.</em></p> <p>So you just need $Y$ to be regular and $X$ to be Cohen-Macaulay (Matsumura actually has a more general statement which also works in the non-finite case).</p> <p>On the other hand, if $Y$ is <em>not</em> regular then $X/Y$ is very rarely flat. Indeed, since the map is finite, flat is the same as $f_* O_Y$ being a locally free $O_X$-module and we can ask how many (local) $O_X$-summands $f_* O_Y$ has as an $O_X$-module.</p> <p><strong>Fact</strong> If $R \subseteq S$ is a finite extension of normal local rings with the same residue field which is etale in codimension 1 (ie, the sort of group quotient that Sasha wrote above), then $S$ has at most one $R$-summand as an $R$-module (in particular, it is nowhere near flat). See the question I asked earlier: <a href="https://mathoverflow.net/questions/236048/number-of-free-summands-of-finite-local-extensions">Number of free summands of finite local extensions</a></p>
modulus-related analytic functions
258,874
7
https://mathoverflow.net/questions/258874/modulus-related-analytic-functions
<p>Let $D\subset\mathbb{C}$ be the open unit disk. Suppose $f,g,F,G:D\rightarrow\mathbb{C}$ are analytic functions linked by $$\vert f(z)\vert^2+\vert g(z)\vert^2=\vert F(z)\vert^2+\vert G(z)\vert^2; \qquad \forall z\in D.$$</p> <blockquote> <p><strong>Question 1.</strong> If $f\neq \alpha g$ and $g\neq\beta f$ for any $\alpha, \beta\in\mathbb{C}$ then is the same true for $F$ and $G$?</p> </blockquote> <p><strong>Caveat.</strong> Not true for real analytic functions: take $f=\sin x, \,g=\cos x, \, F=\frac1{\sqrt{2}}=G$.</p> <p><strong>EDIT.</strong> Encouraged by Christian's positive answer, let's upgrade the problem.</p> <blockquote> <p><strong>Question 2.</strong> Suppose $f, g, h$ are linearly independent (over $\mathbb{C}$) such that $$\vert f\vert^2+\vert g\vert^2+\vert h\vert^2=\vert F\vert^2+\vert G\vert^2+\vert H\vert^2.$$ Should $F, G, H$ be linearly independent, too?</p> </blockquote>
[ { "answer_id": 258877, "score": 8, "text": "<p>Yes, for <strong>Question 1.</strong> </p>\n\n<p>We can divide through by one of the functions on the left-hand side and also assume $F=G$, and then the question becomes: Can $1+|h(z)|^2$ be the square of the absolute value of a holomorphic function, for a non-constant holomorphic $h$?</p>\n\n<p>For this to happen, we need $\\log (1+|h(z)|^2)$ to be harmonic.\nIf you now just work out the Laplacian of this function and use the Cauchy-Riemann equations for $h$, you'll find that the numerator of the expression you obtain from two applications of the quotient rule equals $4((h_1)_x^2+(h_2)_x^2)$, so $h$ has to be constant.</p>\n" }, { "answer_id": 258982, "score": 2, "text": "<p>In fact, more is true. If $f_j$ are linearly independent and $F_i$ are\nlinearly independent, and\n$$\\sum_{j=1}^n|f_j|^2=\\sum_{i=1}^m|F_i|^2,$$\nthen $m=n$ and $F_i$ are obtained from $g_i$ by a unitary transformation.\nSee, for example, <a href=\"https://arxiv.org/pdf/math/0007030.pdf\" rel=\"nofollow noreferrer\">https://arxiv.org/pdf/math/0007030.pdf</a>, section 3. This is called the \"Calabi rigidity\".</p>\n" }, { "answer_id": 258877, "score": 8, "text": "<p>Yes, for <strong>Question 1.</strong> </p>\n\n<p>We can divide through by one of the functions on the left-hand side and also assume $F=G$, and then the question becomes: Can $1+|h(z)|^2$ be the square of the absolute value of a holomorphic function, for a non-constant holomorphic $h$?</p>\n\n<p>For this to happen, we need $\\log (1+|h(z)|^2)$ to be harmonic.\nIf you now just work out the Laplacian of this function and use the Cauchy-Riemann equations for $h$, you'll find that the numerator of the expression you obtain from two applications of the quotient rule equals $4((h_1)_x^2+(h_2)_x^2)$, so $h$ has to be constant.</p>\n" }, { "answer_id": 258982, "score": 2, "text": "<p>In fact, more is true. If $f_j$ are linearly independent and $F_i$ are\nlinearly independent, and\n$$\\sum_{j=1}^n|f_j|^2=\\sum_{i=1}^m|F_i|^2,$$\nthen $m=n$ and $F_i$ are obtained from $g_i$ by a unitary transformation.\nSee, for example, <a href=\"https://arxiv.org/pdf/math/0007030.pdf\" rel=\"nofollow noreferrer\">https://arxiv.org/pdf/math/0007030.pdf</a>, section 3. This is called the \"Calabi rigidity\".</p>\n" } ]
Question: modulus-related analytic functions Answer 1: <p>Yes, for <strong>Question 1.</strong> </p> <p>We can divide through by one of the functions on the left-hand side and also assume $F=G$, and then the question becomes: Can $1+|h(z)|^2$ be the square of the absolute value of a holomorphic function, for a non-constant holomorphic $h$?</p> <p>For this to happen, we need $\log (1+|h(z)|^2)$ to be harmonic. If you now just work out the Laplacian of this function and use the Cauchy-Riemann equations for $h$, you'll find that the numerator of the expression you obtain from two applications of the quotient rule equals $4((h_1)_x^2+(h_2)_x^2)$, so $h$ has to be constant.</p> Answer 2: <p>In fact, more is true. If $f_j$ are linearly independent and $F_i$ are linearly independent, and $$\sum_{j=1}^n|f_j|^2=\sum_{i=1}^m|F_i|^2,$$ then $m=n$ and $F_i$ are obtained from $g_i$ by a unitary transformation. See, for example, <a href="https://arxiv.org/pdf/math/0007030.pdf" rel="nofollow noreferrer">https://arxiv.org/pdf/math/0007030.pdf</a>, section 3. This is called the "Calabi rigidity".</p> Answer 3: <p>Yes, for <strong>Question 1.</strong> </p> <p>We can divide through by one of the functions on the left-hand side and also assume $F=G$, and then the question becomes: Can $1+|h(z)|^2$ be the square of the absolute value of a holomorphic function, for a non-constant holomorphic $h$?</p> <p>For this to happen, we need $\log (1+|h(z)|^2)$ to be harmonic. If you now just work out the Laplacian of this function and use the Cauchy-Riemann equations for $h$, you'll find that the numerator of the expression you obtain from two applications of the quotient rule equals $4((h_1)_x^2+(h_2)_x^2)$, so $h$ has to be constant.</p> Answer 4: <p>In fact, more is true. If $f_j$ are linearly independent and $F_i$ are linearly independent, and $$\sum_{j=1}^n|f_j|^2=\sum_{i=1}^m|F_i|^2,$$ then $m=n$ and $F_i$ are obtained from $g_i$ by a unitary transformation. See, for example, <a href="https://arxiv.org/pdf/math/0007030.pdf" rel="nofollow noreferrer">https://arxiv.org/pdf/math/0007030.pdf</a>, section 3. This is called the "Calabi rigidity".</p>
On the consistency of the definition of the conductor for automorphic forms
259,920
7
https://mathoverflow.net/questions/259920/on-the-consistency-of-the-definition-of-the-conductor-for-automorphic-forms
<p>Let <span class="math-container">$\pi$</span> be an irreducible admissible representation of <span class="math-container">$\mathrm{GL}_2(F)$</span>, where <span class="math-container">$F$</span> is local non-archimedean. The local conductor associated to <span class="math-container">$\pi$</span> can be defined in two usual manners:</p> <p><strong>By its associated L-function</strong></p> <p>Godement and Jacquet associate to it the automorphic L-function <span class="math-container">$L(s, \pi)$</span>. This L-function satisfies a functional equation of the form <span class="math-container">$$L(1-s, \tilde{\pi}) = \varepsilon(s, \pi) L(s, \pi)$$</span></p> <p>This <span class="math-container">$\varepsilon$</span>-factor is essentially of the form <span class="math-container">$c^{s - \frac{1}{2}}$</span>, and we define the appearing real number <span class="math-container">$c(\pi)=c$</span> to be the conductor of <span class="math-container">$\pi$</span>.</p> <p><strong>By some &quot;depth&quot; property</strong></p> <p>The other way to define the conductor is to follow the work of Casselman for <span class="math-container">$\mathrm{GL}_2$</span>, and more generaly JPSS for <span class="math-container">$\mathrm{GL}_n$</span>. We consider the decreasing sequence of compact open congruence subgroups, for <span class="math-container">$r \geqslant 0$</span>: <span class="math-container">$$ K_{0, \mathfrak{p}}\left(\mathfrak{p}^r\right) = \left\{ g \in \mathrm{GL}_2\left(\mathcal{O}_{\mathfrak{p}}\right) \ : \ g \equiv \left( \begin{array}{cc} \star &amp; \star \\ 0 &amp; \star \end{array} \right) \mod \mathfrak{p}^r \right\} \subseteq \mathrm{GL}_2(F). $$</span></p> <p>The conductor of an irreducible admissible infinite-dimensional representation <span class="math-container">$\pi_\mathfrak{p}$</span> of <span class="math-container">$\tilde{G_\mathfrak{p}}$</span> with trivial central character (to ease notations) is then defined by: <span class="math-container">$$ c(\pi) = N\mathfrak{p}^{f(\pi_{\mathfrak{p}})} , $$</span></p> <p>where: <span class="math-container">$$ f\left(\pi_{\mathfrak{p}}\right) = \min \left\{r \geqslant 0 \ : \ \pi_{\mathfrak{p}}^{K_{0, \mathfrak{p}}\left(\mathfrak{p}^r\right)} \neq 0\right\} $$</span></p> <p><strong>Are they consistent?</strong></p> <p>Here is my question: is it obvious that those two definitions are the same?</p>
[ { "answer_id": 259940, "score": 12, "text": "<p>These definitions are consistent, though it's not immediate.</p>\n<p>The conductor quantifies the extent to which <span class=\"math-container\">$\\pi$</span> is ramified. As an aside, I prefer to write <span class=\"math-container\">$c(\\pi)$</span> for the conductor exponent of <span class=\"math-container\">$\\pi$</span>, which is a nonnegative integer, so that <span class=\"math-container\">$\\mathfrak{p}^{c(\\pi)}$</span> is the conductor of <span class=\"math-container\">$\\pi$</span>, and <span class=\"math-container\">$q^{c(\\pi)}$</span> is the absolute conductor of <span class=\"math-container\">$\\pi$</span>, where <span class=\"math-container\">$q = N(\\mathfrak{p}) = \\# \\mathcal{O}_F / \\mathfrak{p}$</span>. This isn't standard terminology though, as these are all unfortunately called the same thing by different people.</p>\n<p>The conductor exponent is tied to a distinguished vector in <span class=\"math-container\">$\\pi$</span>, called the newform of <span class=\"math-container\">$\\pi$</span>. (It is also called the Whittaker newform, the essential Whittaker function, the newvector, or the essential vector; the terminology here hasn't reached a consensus, but newform is best in my opinion because for <span class=\"math-container\">$\\mathrm{GL}_2$</span>, it is the local component of a classical newform.) There are two definitions of the newform of <span class=\"math-container\">$\\pi$</span>, which are equivalent. I'll state them for <span class=\"math-container\">$\\mathrm{GL}_n$</span> instead of <span class=\"math-container\">$\\mathrm{GL}_2$</span>. Throughout, <span class=\"math-container\">$n \\geq 2$</span> and <span class=\"math-container\">$F$</span> is a nonarchimedean local field.</p>\n<ol>\n<li>Recall that for a spherical (that is, unramified principal series) representation <span class=\"math-container\">$\\pi'$</span> of <span class=\"math-container\">$\\mathrm{GL}_{n-1}(F)$</span>, there exists a spherical vector <span class=\"math-container\">$W'^{\\circ}$</span> in the Whittaker model <span class=\"math-container\">$\\mathcal{W}(\\pi',\\overline{\\psi})$</span> of <span class=\"math-container\">$\\pi'$</span>, so that <span class=\"math-container\">$W'^{\\circ}(1_{n-1}) = 1$</span> and <span class=\"math-container\">$W'^{\\circ}(gk) = W(g)$</span> for all <span class=\"math-container\">$g \\in \\mathrm{GL}_{n-1}(F)$</span> and <span class=\"math-container\">$k \\in K_{n-1} = \\mathrm{GL}_{n-1}(\\mathcal{O}_F)$</span>; that such a vector exists and is unique is classical. Now let <span class=\"math-container\">$\\pi$</span> be a (possibly ramified) generic irreducible admissible representation of <span class=\"math-container\">$\\mathrm{GL}_n(F)$</span>. Then there exists a unique Whittaker function <span class=\"math-container\">$W^{\\circ}$</span> in the Whittaker model <span class=\"math-container\">$\\mathcal{W}(\\pi,\\psi)$</span> of <span class=\"math-container\">$\\pi$</span> satisfying\n<span class=\"math-container\">$$W^{\\circ} \\left(g \\begin{pmatrix} k &amp; 0 \\\\ 0 &amp; 1 \\end{pmatrix}\\right) = W^{\\circ}(g), \\qquad W^{\\circ}(1_n) = 1$$</span>\nfor every <span class=\"math-container\">$g \\in \\mathrm{GL}_n(F)$</span> and <span class=\"math-container\">$k \\in K_{n-1}$</span> such that for every spherical representation <span class=\"math-container\">$\\pi'$</span> of <span class=\"math-container\">$\\mathrm{GL}_{n-1}(F)$</span> with associated spherical vector <span class=\"math-container\">$W'^{\\circ}$</span>, the Eulerian integral\n<span class=\"math-container\">$$\\Psi(s,W^{\\circ},W'^{\\circ}) = \\int\\limits_{N_{n-1}(F) \\backslash \\mathrm{GL}_{n-1}(F)} W^{\\circ} \\begin{pmatrix} g &amp; 0 \\\\ 0 &amp; 1 \\end{pmatrix} W'^{\\circ}(g) |\\det g|^{s - 1/2} \\, dg$$</span>\nis equal to the Rankin-Selberg <span class=\"math-container\">$L$</span>-function <span class=\"math-container\">$L(s, \\pi \\times \\pi')$</span>. <span class=\"math-container\">$W^{\\circ}$</span> is called the newform of <span class=\"math-container\">$\\pi$</span>. (In general, <span class=\"math-container\">$\\Psi(s,W,W')/L(s,\\pi \\times \\pi')$</span> is a polynomial in <span class=\"math-container\">$q^{-s}$</span>; the point is that there exists a distinguished choice of <span class=\"math-container\">$W \\in \\mathcal{W}(\\pi,\\psi)$</span>, <span class=\"math-container\">$W' \\in \\mathcal{W}(\\pi',\\overline{\\psi})$</span> for which this polynomial is equal to <span class=\"math-container\">$1$</span>.)</li>\n<li>For a nonnegative integer <span class=\"math-container\">$m$</span>, let <span class=\"math-container\">$K_1(\\mathfrak{p}^m)$</span> denote the congruence subgroup of <span class=\"math-container\">$K_n = \\mathrm{GL}_n(\\mathcal{O}_F)$</span> given by\n<span class=\"math-container\">$$\\left\\{\\begin{pmatrix} a &amp; b \\\\ c &amp; d \\end{pmatrix} \\in K_n : c \\in \\mathrm{Mat}_{1 \\times (n-1)}(\\mathfrak{p}^m), \\\\ d - 1 \\in \\mathfrak{p}^m \\right\\}.$$</span>\n(Note that necessarily <span class=\"math-container\">$a \\in K_{n-1}$</span> and <span class=\"math-container\">$b \\in \\mathrm{Mat}_{n-1 \\times 1}(\\mathcal{O}_F)$</span>.) Let <span class=\"math-container\">$\\pi$</span> be a (possibly ramified) generic irreducible admissible representation of <span class=\"math-container\">$\\mathrm{GL}_n(F)$</span>. Then there exists a minimal <span class=\"math-container\">$m$</span> for which the vector subspace\n<span class=\"math-container\">$$\\pi^{K_1(\\mathfrak{p}^m)} = \\left\\{v \\in \\pi : \\pi(k) \\cdot v = v \\quad \\text{for all $k \\in K_1(\\mathfrak{p}^m)$}\\right\\}$$</span>\nof <span class=\"math-container\">$\\pi$</span> is not equal to <span class=\"math-container\">$\\{0\\}$</span>. We denote by <span class=\"math-container\">$c(\\pi)$</span> this minimal <span class=\"math-container\">$m$</span>. Then <span class=\"math-container\">$\\pi^{K_1(\\mathfrak{p}^{c(\\pi)})}$</span> is one dimensional, so that there exists a unique Whittaker function <span class=\"math-container\">$W^{\\circ}$</span> in the Whittaker model <span class=\"math-container\">$\\mathcal{W}(\\pi,\\psi)$</span> of <span class=\"math-container\">$\\pi$</span> satisfying\n<span class=\"math-container\">$$W^{\\circ} (gk) = W^{\\circ}(g), \\qquad W^{\\circ}(1_n) = 1$$</span>\nfor every <span class=\"math-container\">$g \\in \\mathrm{GL}_n(F)$</span> and <span class=\"math-container\">$k \\in K_1(\\mathfrak{p}^{c(\\pi)})$</span>.</li>\n</ol>\n<p>The original proof of the existence and uniqueness of the newform <span class=\"math-container\">$W^{\\circ}$</span> of <span class=\"math-container\">$\\pi$</span> is via (1), in the paper <a href=\"https://eudml.org/doc/163506\" rel=\"nofollow noreferrer\">Conducteur des répresentations du group linéaire</a> by Jacquet, Piatetski-Shapiro, and Shalika. However, it was noticed by Matringe several years ago that the proof was in fact incomplete. He gave a <a href=\"http://www.math.uiuc.edu/documenta/vol-18/37.pdf\" rel=\"nofollow noreferrer\">correct proof</a>, as did <a href=\"http://msp.org/pjm/2012/260-2/pjm-v260-n2-p09-p.pdf\" rel=\"nofollow noreferrer\">Jacquet</a>.</p>\n<p>In the last section of Jacquet, Piatetski-Shapiro, and Shalika's paper, they show that (1) implies (2) via the functional equation for <span class=\"math-container\">$L(s,\\pi \\times \\pi')$</span>. They make use of the fact that the epsilon factor <span class=\"math-container\">$\\epsilon(s,\\pi \\times \\pi',\\psi)$</span> is equal to <span class=\"math-container\">$\\epsilon(1/2, \\pi \\times \\pi',\\psi) q^{-(n - 1) c(\\pi) (s - 1/2)}$</span> for some nonnegative integer <span class=\"math-container\">$c(\\pi)$</span>, and it is precisely this nonnegative integer <span class=\"math-container\">$c(\\pi)$</span> that satisfies <span class=\"math-container\">$\\pi^{K_1(\\mathfrak{p}^{c(\\pi)})} \\ni W^{\\circ}$</span> and <span class=\"math-container\">$\\pi^{K_1(\\mathfrak{p}^m)} = \\{0\\}$</span> for all <span class=\"math-container\">$m &lt; c(\\pi)$</span>.</p>\n<p>I don't know if the fact that (2) implies (1) has appeared anywhere in print, but I do know how to prove it. There is a <a href=\"https://projecteuclid.org/euclid.jmsj/1390600834\" rel=\"nofollow noreferrer\">paper</a> by Miyauchi that shows that the (Whittaker) newform <span class=\"math-container\">$W^{\\circ}$</span> given by (2) is such that\n<span class=\"math-container\">$$\\Psi(s,W^{\\circ}) = \\int_{F^{\\times}} W^{\\circ} \\begin{pmatrix} x &amp; 0 &amp; \\cdots &amp; 0 \\\\ 0 &amp; 1 &amp; \\cdots &amp; 0 \\\\ \\vdots &amp; \\vdots &amp; \\ddots &amp; \\vdots \\\\ 0 &amp; &amp; \\cdots &amp; 1 \\end{pmatrix} |x|^{s - \\frac{n - 1}{2}} \\, d^{\\times} x$$</span>\nis equal to the <span class=\"math-container\">$L$</span>-function <span class=\"math-container\">$L(s,\\pi)$</span>, and the same method of proof (using Hecke operators) shows that <span class=\"math-container\">$\\Psi(s,W^{\\circ},W'^{\\circ}) = L(s, \\pi \\times \\pi')$</span> for all spherical representations <span class=\"math-container\">$\\pi'$</span> of <span class=\"math-container\">$\\mathrm{GL}_{n-1}(F)$</span>. Now you can work backwards to show that <span class=\"math-container\">$\\epsilon(s,\\pi \\times \\pi',\\psi)$</span> is equal to <span class=\"math-container\">$\\epsilon(1/2, \\pi \\times \\pi',\\psi) q^{-(n - 1) c(\\pi) (s - 1/2)}$</span>, as in Martin Dickson's answer.</p>\n" }, { "answer_id": 259938, "score": 10, "text": "<p>Yes this is true, but I don't think it's completely obvious. The following argument is taken from Roberts--Schmidt ``Local newforms for $GSp_4$'', which uses the same argument as the $GL_2$ case.</p>\n\n<p>Jacquet--Langlands proves the local functional equation\n$$\nZ(1-s, \\pi(w_0) W) = \\gamma(s, W) Z(s, W)\n$$\nwhere $W$ is in the Whittaker model, $Z(s, W)$ is the usual zeta integral, and $w_0 = \\begin{bmatrix} &amp; -1 \\\\ 1 \\end{bmatrix}$ is the long Weyl element. For any positive integer $n$, define $w_n = \\begin{bmatrix} &amp; -1 \\\\ \\varpi^n \\end{bmatrix}$; we can write the above as\n$$\nZ(1-s, \\pi(w_n) W) = q^{n(s-1/2)} \\gamma(s, W) Z(s, W).\n$$</p>\n\n<p>Now take $\\pi$ of conductor $\\mathfrak{p}^n$ (in the depth sense), and $W$ a generator for the space of $K(\\mathfrak{p}^n)$-fixed vectors. It is known (presumably implicit in J--L) that one can choose $W$ s.t. $Z(s, W) = L(s, \\pi)$. Also, one easily sees that $\\pi(w_n)$ normalises $K(\\mathfrak{p}^n)$ and hence acts on its fixed vectors, and so the LHS is also the corresponding local $L$-function, times the eigenvalue of $w_n$. Now dividing by the local $L$-functions to get the definition of the $\\epsilon$-factor, one now sees that the conductor in the $\\epsilon$-factor is what it should be.</p>\n" }, { "answer_id": 259940, "score": 12, "text": "<p>These definitions are consistent, though it's not immediate.</p>\n<p>The conductor quantifies the extent to which <span class=\"math-container\">$\\pi$</span> is ramified. As an aside, I prefer to write <span class=\"math-container\">$c(\\pi)$</span> for the conductor exponent of <span class=\"math-container\">$\\pi$</span>, which is a nonnegative integer, so that <span class=\"math-container\">$\\mathfrak{p}^{c(\\pi)}$</span> is the conductor of <span class=\"math-container\">$\\pi$</span>, and <span class=\"math-container\">$q^{c(\\pi)}$</span> is the absolute conductor of <span class=\"math-container\">$\\pi$</span>, where <span class=\"math-container\">$q = N(\\mathfrak{p}) = \\# \\mathcal{O}_F / \\mathfrak{p}$</span>. This isn't standard terminology though, as these are all unfortunately called the same thing by different people.</p>\n<p>The conductor exponent is tied to a distinguished vector in <span class=\"math-container\">$\\pi$</span>, called the newform of <span class=\"math-container\">$\\pi$</span>. (It is also called the Whittaker newform, the essential Whittaker function, the newvector, or the essential vector; the terminology here hasn't reached a consensus, but newform is best in my opinion because for <span class=\"math-container\">$\\mathrm{GL}_2$</span>, it is the local component of a classical newform.) There are two definitions of the newform of <span class=\"math-container\">$\\pi$</span>, which are equivalent. I'll state them for <span class=\"math-container\">$\\mathrm{GL}_n$</span> instead of <span class=\"math-container\">$\\mathrm{GL}_2$</span>. Throughout, <span class=\"math-container\">$n \\geq 2$</span> and <span class=\"math-container\">$F$</span> is a nonarchimedean local field.</p>\n<ol>\n<li>Recall that for a spherical (that is, unramified principal series) representation <span class=\"math-container\">$\\pi'$</span> of <span class=\"math-container\">$\\mathrm{GL}_{n-1}(F)$</span>, there exists a spherical vector <span class=\"math-container\">$W'^{\\circ}$</span> in the Whittaker model <span class=\"math-container\">$\\mathcal{W}(\\pi',\\overline{\\psi})$</span> of <span class=\"math-container\">$\\pi'$</span>, so that <span class=\"math-container\">$W'^{\\circ}(1_{n-1}) = 1$</span> and <span class=\"math-container\">$W'^{\\circ}(gk) = W(g)$</span> for all <span class=\"math-container\">$g \\in \\mathrm{GL}_{n-1}(F)$</span> and <span class=\"math-container\">$k \\in K_{n-1} = \\mathrm{GL}_{n-1}(\\mathcal{O}_F)$</span>; that such a vector exists and is unique is classical. Now let <span class=\"math-container\">$\\pi$</span> be a (possibly ramified) generic irreducible admissible representation of <span class=\"math-container\">$\\mathrm{GL}_n(F)$</span>. Then there exists a unique Whittaker function <span class=\"math-container\">$W^{\\circ}$</span> in the Whittaker model <span class=\"math-container\">$\\mathcal{W}(\\pi,\\psi)$</span> of <span class=\"math-container\">$\\pi$</span> satisfying\n<span class=\"math-container\">$$W^{\\circ} \\left(g \\begin{pmatrix} k &amp; 0 \\\\ 0 &amp; 1 \\end{pmatrix}\\right) = W^{\\circ}(g), \\qquad W^{\\circ}(1_n) = 1$$</span>\nfor every <span class=\"math-container\">$g \\in \\mathrm{GL}_n(F)$</span> and <span class=\"math-container\">$k \\in K_{n-1}$</span> such that for every spherical representation <span class=\"math-container\">$\\pi'$</span> of <span class=\"math-container\">$\\mathrm{GL}_{n-1}(F)$</span> with associated spherical vector <span class=\"math-container\">$W'^{\\circ}$</span>, the Eulerian integral\n<span class=\"math-container\">$$\\Psi(s,W^{\\circ},W'^{\\circ}) = \\int\\limits_{N_{n-1}(F) \\backslash \\mathrm{GL}_{n-1}(F)} W^{\\circ} \\begin{pmatrix} g &amp; 0 \\\\ 0 &amp; 1 \\end{pmatrix} W'^{\\circ}(g) |\\det g|^{s - 1/2} \\, dg$$</span>\nis equal to the Rankin-Selberg <span class=\"math-container\">$L$</span>-function <span class=\"math-container\">$L(s, \\pi \\times \\pi')$</span>. <span class=\"math-container\">$W^{\\circ}$</span> is called the newform of <span class=\"math-container\">$\\pi$</span>. (In general, <span class=\"math-container\">$\\Psi(s,W,W')/L(s,\\pi \\times \\pi')$</span> is a polynomial in <span class=\"math-container\">$q^{-s}$</span>; the point is that there exists a distinguished choice of <span class=\"math-container\">$W \\in \\mathcal{W}(\\pi,\\psi)$</span>, <span class=\"math-container\">$W' \\in \\mathcal{W}(\\pi',\\overline{\\psi})$</span> for which this polynomial is equal to <span class=\"math-container\">$1$</span>.)</li>\n<li>For a nonnegative integer <span class=\"math-container\">$m$</span>, let <span class=\"math-container\">$K_1(\\mathfrak{p}^m)$</span> denote the congruence subgroup of <span class=\"math-container\">$K_n = \\mathrm{GL}_n(\\mathcal{O}_F)$</span> given by\n<span class=\"math-container\">$$\\left\\{\\begin{pmatrix} a &amp; b \\\\ c &amp; d \\end{pmatrix} \\in K_n : c \\in \\mathrm{Mat}_{1 \\times (n-1)}(\\mathfrak{p}^m), \\\\ d - 1 \\in \\mathfrak{p}^m \\right\\}.$$</span>\n(Note that necessarily <span class=\"math-container\">$a \\in K_{n-1}$</span> and <span class=\"math-container\">$b \\in \\mathrm{Mat}_{n-1 \\times 1}(\\mathcal{O}_F)$</span>.) Let <span class=\"math-container\">$\\pi$</span> be a (possibly ramified) generic irreducible admissible representation of <span class=\"math-container\">$\\mathrm{GL}_n(F)$</span>. Then there exists a minimal <span class=\"math-container\">$m$</span> for which the vector subspace\n<span class=\"math-container\">$$\\pi^{K_1(\\mathfrak{p}^m)} = \\left\\{v \\in \\pi : \\pi(k) \\cdot v = v \\quad \\text{for all $k \\in K_1(\\mathfrak{p}^m)$}\\right\\}$$</span>\nof <span class=\"math-container\">$\\pi$</span> is not equal to <span class=\"math-container\">$\\{0\\}$</span>. We denote by <span class=\"math-container\">$c(\\pi)$</span> this minimal <span class=\"math-container\">$m$</span>. Then <span class=\"math-container\">$\\pi^{K_1(\\mathfrak{p}^{c(\\pi)})}$</span> is one dimensional, so that there exists a unique Whittaker function <span class=\"math-container\">$W^{\\circ}$</span> in the Whittaker model <span class=\"math-container\">$\\mathcal{W}(\\pi,\\psi)$</span> of <span class=\"math-container\">$\\pi$</span> satisfying\n<span class=\"math-container\">$$W^{\\circ} (gk) = W^{\\circ}(g), \\qquad W^{\\circ}(1_n) = 1$$</span>\nfor every <span class=\"math-container\">$g \\in \\mathrm{GL}_n(F)$</span> and <span class=\"math-container\">$k \\in K_1(\\mathfrak{p}^{c(\\pi)})$</span>.</li>\n</ol>\n<p>The original proof of the existence and uniqueness of the newform <span class=\"math-container\">$W^{\\circ}$</span> of <span class=\"math-container\">$\\pi$</span> is via (1), in the paper <a href=\"https://eudml.org/doc/163506\" rel=\"nofollow noreferrer\">Conducteur des répresentations du group linéaire</a> by Jacquet, Piatetski-Shapiro, and Shalika. However, it was noticed by Matringe several years ago that the proof was in fact incomplete. He gave a <a href=\"http://www.math.uiuc.edu/documenta/vol-18/37.pdf\" rel=\"nofollow noreferrer\">correct proof</a>, as did <a href=\"http://msp.org/pjm/2012/260-2/pjm-v260-n2-p09-p.pdf\" rel=\"nofollow noreferrer\">Jacquet</a>.</p>\n<p>In the last section of Jacquet, Piatetski-Shapiro, and Shalika's paper, they show that (1) implies (2) via the functional equation for <span class=\"math-container\">$L(s,\\pi \\times \\pi')$</span>. They make use of the fact that the epsilon factor <span class=\"math-container\">$\\epsilon(s,\\pi \\times \\pi',\\psi)$</span> is equal to <span class=\"math-container\">$\\epsilon(1/2, \\pi \\times \\pi',\\psi) q^{-(n - 1) c(\\pi) (s - 1/2)}$</span> for some nonnegative integer <span class=\"math-container\">$c(\\pi)$</span>, and it is precisely this nonnegative integer <span class=\"math-container\">$c(\\pi)$</span> that satisfies <span class=\"math-container\">$\\pi^{K_1(\\mathfrak{p}^{c(\\pi)})} \\ni W^{\\circ}$</span> and <span class=\"math-container\">$\\pi^{K_1(\\mathfrak{p}^m)} = \\{0\\}$</span> for all <span class=\"math-container\">$m &lt; c(\\pi)$</span>.</p>\n<p>I don't know if the fact that (2) implies (1) has appeared anywhere in print, but I do know how to prove it. There is a <a href=\"https://projecteuclid.org/euclid.jmsj/1390600834\" rel=\"nofollow noreferrer\">paper</a> by Miyauchi that shows that the (Whittaker) newform <span class=\"math-container\">$W^{\\circ}$</span> given by (2) is such that\n<span class=\"math-container\">$$\\Psi(s,W^{\\circ}) = \\int_{F^{\\times}} W^{\\circ} \\begin{pmatrix} x &amp; 0 &amp; \\cdots &amp; 0 \\\\ 0 &amp; 1 &amp; \\cdots &amp; 0 \\\\ \\vdots &amp; \\vdots &amp; \\ddots &amp; \\vdots \\\\ 0 &amp; &amp; \\cdots &amp; 1 \\end{pmatrix} |x|^{s - \\frac{n - 1}{2}} \\, d^{\\times} x$$</span>\nis equal to the <span class=\"math-container\">$L$</span>-function <span class=\"math-container\">$L(s,\\pi)$</span>, and the same method of proof (using Hecke operators) shows that <span class=\"math-container\">$\\Psi(s,W^{\\circ},W'^{\\circ}) = L(s, \\pi \\times \\pi')$</span> for all spherical representations <span class=\"math-container\">$\\pi'$</span> of <span class=\"math-container\">$\\mathrm{GL}_{n-1}(F)$</span>. Now you can work backwards to show that <span class=\"math-container\">$\\epsilon(s,\\pi \\times \\pi',\\psi)$</span> is equal to <span class=\"math-container\">$\\epsilon(1/2, \\pi \\times \\pi',\\psi) q^{-(n - 1) c(\\pi) (s - 1/2)}$</span>, as in Martin Dickson's answer.</p>\n" }, { "answer_id": 259938, "score": 10, "text": "<p>Yes this is true, but I don't think it's completely obvious. The following argument is taken from Roberts--Schmidt ``Local newforms for $GSp_4$'', which uses the same argument as the $GL_2$ case.</p>\n\n<p>Jacquet--Langlands proves the local functional equation\n$$\nZ(1-s, \\pi(w_0) W) = \\gamma(s, W) Z(s, W)\n$$\nwhere $W$ is in the Whittaker model, $Z(s, W)$ is the usual zeta integral, and $w_0 = \\begin{bmatrix} &amp; -1 \\\\ 1 \\end{bmatrix}$ is the long Weyl element. For any positive integer $n$, define $w_n = \\begin{bmatrix} &amp; -1 \\\\ \\varpi^n \\end{bmatrix}$; we can write the above as\n$$\nZ(1-s, \\pi(w_n) W) = q^{n(s-1/2)} \\gamma(s, W) Z(s, W).\n$$</p>\n\n<p>Now take $\\pi$ of conductor $\\mathfrak{p}^n$ (in the depth sense), and $W$ a generator for the space of $K(\\mathfrak{p}^n)$-fixed vectors. It is known (presumably implicit in J--L) that one can choose $W$ s.t. $Z(s, W) = L(s, \\pi)$. Also, one easily sees that $\\pi(w_n)$ normalises $K(\\mathfrak{p}^n)$ and hence acts on its fixed vectors, and so the LHS is also the corresponding local $L$-function, times the eigenvalue of $w_n$. Now dividing by the local $L$-functions to get the definition of the $\\epsilon$-factor, one now sees that the conductor in the $\\epsilon$-factor is what it should be.</p>\n" } ]
Question: On the consistency of the definition of the conductor for automorphic forms Answer 1: <p>These definitions are consistent, though it's not immediate.</p> <p>The conductor quantifies the extent to which <span class="math-container">$\pi$</span> is ramified. As an aside, I prefer to write <span class="math-container">$c(\pi)$</span> for the conductor exponent of <span class="math-container">$\pi$</span>, which is a nonnegative integer, so that <span class="math-container">$\mathfrak{p}^{c(\pi)}$</span> is the conductor of <span class="math-container">$\pi$</span>, and <span class="math-container">$q^{c(\pi)}$</span> is the absolute conductor of <span class="math-container">$\pi$</span>, where <span class="math-container">$q = N(\mathfrak{p}) = \# \mathcal{O}_F / \mathfrak{p}$</span>. This isn't standard terminology though, as these are all unfortunately called the same thing by different people.</p> <p>The conductor exponent is tied to a distinguished vector in <span class="math-container">$\pi$</span>, called the newform of <span class="math-container">$\pi$</span>. (It is also called the Whittaker newform, the essential Whittaker function, the newvector, or the essential vector; the terminology here hasn't reached a consensus, but newform is best in my opinion because for <span class="math-container">$\mathrm{GL}_2$</span>, it is the local component of a classical newform.) There are two definitions of the newform of <span class="math-container">$\pi$</span>, which are equivalent. I'll state them for <span class="math-container">$\mathrm{GL}_n$</span> instead of <span class="math-container">$\mathrm{GL}_2$</span>. Throughout, <span class="math-container">$n \geq 2$</span> and <span class="math-container">$F$</span> is a nonarchimedean local field.</p> <ol> <li>Recall that for a spherical (that is, unramified principal series) representation <span class="math-container">$\pi'$</span> of <span class="math-container">$\mathrm{GL}_{n-1}(F)$</span>, there exists a spherical vector <span class="math-container">$W'^{\circ}$</span> in the Whittaker model <span class="math-container">$\mathcal{W}(\pi',\overline{\psi})$</span> of <span class="math-container">$\pi'$</span>, so that <span class="math-container">$W'^{\circ}(1_{n-1}) = 1$</span> and <span class="math-container">$W'^{\circ}(gk) = W(g)$</span> for all <span class="math-container">$g \in \mathrm{GL}_{n-1}(F)$</span> and <span class="math-container">$k \in K_{n-1} = \mathrm{GL}_{n-1}(\mathcal{O}_F)$</span>; that such a vector exists and is unique is classical. Now let <span class="math-container">$\pi$</span> be a (possibly ramified) generic irreducible admissible representation of <span class="math-container">$\mathrm{GL}_n(F)$</span>. Then there exists a unique Whittaker function <span class="math-container">$W^{\circ}$</span> in the Whittaker model <span class="math-container">$\mathcal{W}(\pi,\psi)$</span> of <span class="math-container">$\pi$</span> satisfying <span class="math-container">$$W^{\circ} \left(g \begin{pmatrix} k &amp; 0 \\ 0 &amp; 1 \end{pmatrix}\right) = W^{\circ}(g), \qquad W^{\circ}(1_n) = 1$$</span> for every <span class="math-container">$g \in \mathrm{GL}_n(F)$</span> and <span class="math-container">$k \in K_{n-1}$</span> such that for every spherical representation <span class="math-container">$\pi'$</span> of <span class="math-container">$\mathrm{GL}_{n-1}(F)$</span> with associated spherical vector <span class="math-container">$W'^{\circ}$</span>, the Eulerian integral <span class="math-container">$$\Psi(s,W^{\circ},W'^{\circ}) = \int\limits_{N_{n-1}(F) \backslash \mathrm{GL}_{n-1}(F)} W^{\circ} \begin{pmatrix} g &amp; 0 \\ 0 &amp; 1 \end{pmatrix} W'^{\circ}(g) |\det g|^{s - 1/2} \, dg$$</span> is equal to the Rankin-Selberg <span class="math-container">$L$</span>-function <span class="math-container">$L(s, \pi \times \pi')$</span>. <span class="math-container">$W^{\circ}$</span> is called the newform of <span class="math-container">$\pi$</span>. (In general, <span class="math-container">$\Psi(s,W,W')/L(s,\pi \times \pi')$</span> is a polynomial in <span class="math-container">$q^{-s}$</span>; the point is that there exists a distinguished choice of <span class="math-container">$W \in \mathcal{W}(\pi,\psi)$</span>, <span class="math-container">$W' \in \mathcal{W}(\pi',\overline{\psi})$</span> for which this polynomial is equal to <span class="math-container">$1$</span>.)</li> <li>For a nonnegative integer <span class="math-container">$m$</span>, let <span class="math-container">$K_1(\mathfrak{p}^m)$</span> denote the congruence subgroup of <span class="math-container">$K_n = \mathrm{GL}_n(\mathcal{O}_F)$</span> given by <span class="math-container">$$\left\{\begin{pmatrix} a &amp; b \\ c &amp; d \end{pmatrix} \in K_n : c \in \mathrm{Mat}_{1 \times (n-1)}(\mathfrak{p}^m), \\ d - 1 \in \mathfrak{p}^m \right\}.$$</span> (Note that necessarily <span class="math-container">$a \in K_{n-1}$</span> and <span class="math-container">$b \in \mathrm{Mat}_{n-1 \times 1}(\mathcal{O}_F)$</span>.) Let <span class="math-container">$\pi$</span> be a (possibly ramified) generic irreducible admissible representation of <span class="math-container">$\mathrm{GL}_n(F)$</span>. Then there exists a minimal <span class="math-container">$m$</span> for which the vector subspace <span class="math-container">$$\pi^{K_1(\mathfrak{p}^m)} = \left\{v \in \pi : \pi(k) \cdot v = v \quad \text{for all $k \in K_1(\mathfrak{p}^m)$}\right\}$$</span> of <span class="math-container">$\pi$</span> is not equal to <span class="math-container">$\{0\}$</span>. We denote by <span class="math-container">$c(\pi)$</span> this minimal <span class="math-container">$m$</span>. Then <span class="math-container">$\pi^{K_1(\mathfrak{p}^{c(\pi)})}$</span> is one dimensional, so that there exists a unique Whittaker function <span class="math-container">$W^{\circ}$</span> in the Whittaker model <span class="math-container">$\mathcal{W}(\pi,\psi)$</span> of <span class="math-container">$\pi$</span> satisfying <span class="math-container">$$W^{\circ} (gk) = W^{\circ}(g), \qquad W^{\circ}(1_n) = 1$$</span> for every <span class="math-container">$g \in \mathrm{GL}_n(F)$</span> and <span class="math-container">$k \in K_1(\mathfrak{p}^{c(\pi)})$</span>.</li> </ol> <p>The original proof of the existence and uniqueness of the newform <span class="math-container">$W^{\circ}$</span> of <span class="math-container">$\pi$</span> is via (1), in the paper <a href="https://eudml.org/doc/163506" rel="nofollow noreferrer">Conducteur des répresentations du group linéaire</a> by Jacquet, Piatetski-Shapiro, and Shalika. However, it was noticed by Matringe several years ago that the proof was in fact incomplete. He gave a <a href="http://www.math.uiuc.edu/documenta/vol-18/37.pdf" rel="nofollow noreferrer">correct proof</a>, as did <a href="http://msp.org/pjm/2012/260-2/pjm-v260-n2-p09-p.pdf" rel="nofollow noreferrer">Jacquet</a>.</p> <p>In the last section of Jacquet, Piatetski-Shapiro, and Shalika's paper, they show that (1) implies (2) via the functional equation for <span class="math-container">$L(s,\pi \times \pi')$</span>. They make use of the fact that the epsilon factor <span class="math-container">$\epsilon(s,\pi \times \pi',\psi)$</span> is equal to <span class="math-container">$\epsilon(1/2, \pi \times \pi',\psi) q^{-(n - 1) c(\pi) (s - 1/2)}$</span> for some nonnegative integer <span class="math-container">$c(\pi)$</span>, and it is precisely this nonnegative integer <span class="math-container">$c(\pi)$</span> that satisfies <span class="math-container">$\pi^{K_1(\mathfrak{p}^{c(\pi)})} \ni W^{\circ}$</span> and <span class="math-container">$\pi^{K_1(\mathfrak{p}^m)} = \{0\}$</span> for all <span class="math-container">$m &lt; c(\pi)$</span>.</p> <p>I don't know if the fact that (2) implies (1) has appeared anywhere in print, but I do know how to prove it. There is a <a href="https://projecteuclid.org/euclid.jmsj/1390600834" rel="nofollow noreferrer">paper</a> by Miyauchi that shows that the (Whittaker) newform <span class="math-container">$W^{\circ}$</span> given by (2) is such that <span class="math-container">$$\Psi(s,W^{\circ}) = \int_{F^{\times}} W^{\circ} \begin{pmatrix} x &amp; 0 &amp; \cdots &amp; 0 \\ 0 &amp; 1 &amp; \cdots &amp; 0 \\ \vdots &amp; \vdots &amp; \ddots &amp; \vdots \\ 0 &amp; &amp; \cdots &amp; 1 \end{pmatrix} |x|^{s - \frac{n - 1}{2}} \, d^{\times} x$$</span> is equal to the <span class="math-container">$L$</span>-function <span class="math-container">$L(s,\pi)$</span>, and the same method of proof (using Hecke operators) shows that <span class="math-container">$\Psi(s,W^{\circ},W'^{\circ}) = L(s, \pi \times \pi')$</span> for all spherical representations <span class="math-container">$\pi'$</span> of <span class="math-container">$\mathrm{GL}_{n-1}(F)$</span>. Now you can work backwards to show that <span class="math-container">$\epsilon(s,\pi \times \pi',\psi)$</span> is equal to <span class="math-container">$\epsilon(1/2, \pi \times \pi',\psi) q^{-(n - 1) c(\pi) (s - 1/2)}$</span>, as in Martin Dickson's answer.</p> Answer 2: <p>Yes this is true, but I don't think it's completely obvious. The following argument is taken from Roberts--Schmidt ``Local newforms for $GSp_4$'', which uses the same argument as the $GL_2$ case.</p> <p>Jacquet--Langlands proves the local functional equation $$ Z(1-s, \pi(w_0) W) = \gamma(s, W) Z(s, W) $$ where $W$ is in the Whittaker model, $Z(s, W)$ is the usual zeta integral, and $w_0 = \begin{bmatrix} &amp; -1 \\ 1 \end{bmatrix}$ is the long Weyl element. For any positive integer $n$, define $w_n = \begin{bmatrix} &amp; -1 \\ \varpi^n \end{bmatrix}$; we can write the above as $$ Z(1-s, \pi(w_n) W) = q^{n(s-1/2)} \gamma(s, W) Z(s, W). $$</p> <p>Now take $\pi$ of conductor $\mathfrak{p}^n$ (in the depth sense), and $W$ a generator for the space of $K(\mathfrak{p}^n)$-fixed vectors. It is known (presumably implicit in J--L) that one can choose $W$ s.t. $Z(s, W) = L(s, \pi)$. Also, one easily sees that $\pi(w_n)$ normalises $K(\mathfrak{p}^n)$ and hence acts on its fixed vectors, and so the LHS is also the corresponding local $L$-function, times the eigenvalue of $w_n$. Now dividing by the local $L$-functions to get the definition of the $\epsilon$-factor, one now sees that the conductor in the $\epsilon$-factor is what it should be.</p> Answer 3: <p>These definitions are consistent, though it's not immediate.</p> <p>The conductor quantifies the extent to which <span class="math-container">$\pi$</span> is ramified. As an aside, I prefer to write <span class="math-container">$c(\pi)$</span> for the conductor exponent of <span class="math-container">$\pi$</span>, which is a nonnegative integer, so that <span class="math-container">$\mathfrak{p}^{c(\pi)}$</span> is the conductor of <span class="math-container">$\pi$</span>, and <span class="math-container">$q^{c(\pi)}$</span> is the absolute conductor of <span class="math-container">$\pi$</span>, where <span class="math-container">$q = N(\mathfrak{p}) = \# \mathcal{O}_F / \mathfrak{p}$</span>. This isn't standard terminology though, as these are all unfortunately called the same thing by different people.</p> <p>The conductor exponent is tied to a distinguished vector in <span class="math-container">$\pi$</span>, called the newform of <span class="math-container">$\pi$</span>. (It is also called the Whittaker newform, the essential Whittaker function, the newvector, or the essential vector; the terminology here hasn't reached a consensus, but newform is best in my opinion because for <span class="math-container">$\mathrm{GL}_2$</span>, it is the local component of a classical newform.) There are two definitions of the newform of <span class="math-container">$\pi$</span>, which are equivalent. I'll state them for <span class="math-container">$\mathrm{GL}_n$</span> instead of <span class="math-container">$\mathrm{GL}_2$</span>. Throughout, <span class="math-container">$n \geq 2$</span> and <span class="math-container">$F$</span> is a nonarchimedean local field.</p> <ol> <li>Recall that for a spherical (that is, unramified principal series) representation <span class="math-container">$\pi'$</span> of <span class="math-container">$\mathrm{GL}_{n-1}(F)$</span>, there exists a spherical vector <span class="math-container">$W'^{\circ}$</span> in the Whittaker model <span class="math-container">$\mathcal{W}(\pi',\overline{\psi})$</span> of <span class="math-container">$\pi'$</span>, so that <span class="math-container">$W'^{\circ}(1_{n-1}) = 1$</span> and <span class="math-container">$W'^{\circ}(gk) = W(g)$</span> for all <span class="math-container">$g \in \mathrm{GL}_{n-1}(F)$</span> and <span class="math-container">$k \in K_{n-1} = \mathrm{GL}_{n-1}(\mathcal{O}_F)$</span>; that such a vector exists and is unique is classical. Now let <span class="math-container">$\pi$</span> be a (possibly ramified) generic irreducible admissible representation of <span class="math-container">$\mathrm{GL}_n(F)$</span>. Then there exists a unique Whittaker function <span class="math-container">$W^{\circ}$</span> in the Whittaker model <span class="math-container">$\mathcal{W}(\pi,\psi)$</span> of <span class="math-container">$\pi$</span> satisfying <span class="math-container">$$W^{\circ} \left(g \begin{pmatrix} k &amp; 0 \\ 0 &amp; 1 \end{pmatrix}\right) = W^{\circ}(g), \qquad W^{\circ}(1_n) = 1$$</span> for every <span class="math-container">$g \in \mathrm{GL}_n(F)$</span> and <span class="math-container">$k \in K_{n-1}$</span> such that for every spherical representation <span class="math-container">$\pi'$</span> of <span class="math-container">$\mathrm{GL}_{n-1}(F)$</span> with associated spherical vector <span class="math-container">$W'^{\circ}$</span>, the Eulerian integral <span class="math-container">$$\Psi(s,W^{\circ},W'^{\circ}) = \int\limits_{N_{n-1}(F) \backslash \mathrm{GL}_{n-1}(F)} W^{\circ} \begin{pmatrix} g &amp; 0 \\ 0 &amp; 1 \end{pmatrix} W'^{\circ}(g) |\det g|^{s - 1/2} \, dg$$</span> is equal to the Rankin-Selberg <span class="math-container">$L$</span>-function <span class="math-container">$L(s, \pi \times \pi')$</span>. <span class="math-container">$W^{\circ}$</span> is called the newform of <span class="math-container">$\pi$</span>. (In general, <span class="math-container">$\Psi(s,W,W')/L(s,\pi \times \pi')$</span> is a polynomial in <span class="math-container">$q^{-s}$</span>; the point is that there exists a distinguished choice of <span class="math-container">$W \in \mathcal{W}(\pi,\psi)$</span>, <span class="math-container">$W' \in \mathcal{W}(\pi',\overline{\psi})$</span> for which this polynomial is equal to <span class="math-container">$1$</span>.)</li> <li>For a nonnegative integer <span class="math-container">$m$</span>, let <span class="math-container">$K_1(\mathfrak{p}^m)$</span> denote the congruence subgroup of <span class="math-container">$K_n = \mathrm{GL}_n(\mathcal{O}_F)$</span> given by <span class="math-container">$$\left\{\begin{pmatrix} a &amp; b \\ c &amp; d \end{pmatrix} \in K_n : c \in \mathrm{Mat}_{1 \times (n-1)}(\mathfrak{p}^m), \\ d - 1 \in \mathfrak{p}^m \right\}.$$</span> (Note that necessarily <span class="math-container">$a \in K_{n-1}$</span> and <span class="math-container">$b \in \mathrm{Mat}_{n-1 \times 1}(\mathcal{O}_F)$</span>.) Let <span class="math-container">$\pi$</span> be a (possibly ramified) generic irreducible admissible representation of <span class="math-container">$\mathrm{GL}_n(F)$</span>. Then there exists a minimal <span class="math-container">$m$</span> for which the vector subspace <span class="math-container">$$\pi^{K_1(\mathfrak{p}^m)} = \left\{v \in \pi : \pi(k) \cdot v = v \quad \text{for all $k \in K_1(\mathfrak{p}^m)$}\right\}$$</span> of <span class="math-container">$\pi$</span> is not equal to <span class="math-container">$\{0\}$</span>. We denote by <span class="math-container">$c(\pi)$</span> this minimal <span class="math-container">$m$</span>. Then <span class="math-container">$\pi^{K_1(\mathfrak{p}^{c(\pi)})}$</span> is one dimensional, so that there exists a unique Whittaker function <span class="math-container">$W^{\circ}$</span> in the Whittaker model <span class="math-container">$\mathcal{W}(\pi,\psi)$</span> of <span class="math-container">$\pi$</span> satisfying <span class="math-container">$$W^{\circ} (gk) = W^{\circ}(g), \qquad W^{\circ}(1_n) = 1$$</span> for every <span class="math-container">$g \in \mathrm{GL}_n(F)$</span> and <span class="math-container">$k \in K_1(\mathfrak{p}^{c(\pi)})$</span>.</li> </ol> <p>The original proof of the existence and uniqueness of the newform <span class="math-container">$W^{\circ}$</span> of <span class="math-container">$\pi$</span> is via (1), in the paper <a href="https://eudml.org/doc/163506" rel="nofollow noreferrer">Conducteur des répresentations du group linéaire</a> by Jacquet, Piatetski-Shapiro, and Shalika. However, it was noticed by Matringe several years ago that the proof was in fact incomplete. He gave a <a href="http://www.math.uiuc.edu/documenta/vol-18/37.pdf" rel="nofollow noreferrer">correct proof</a>, as did <a href="http://msp.org/pjm/2012/260-2/pjm-v260-n2-p09-p.pdf" rel="nofollow noreferrer">Jacquet</a>.</p> <p>In the last section of Jacquet, Piatetski-Shapiro, and Shalika's paper, they show that (1) implies (2) via the functional equation for <span class="math-container">$L(s,\pi \times \pi')$</span>. They make use of the fact that the epsilon factor <span class="math-container">$\epsilon(s,\pi \times \pi',\psi)$</span> is equal to <span class="math-container">$\epsilon(1/2, \pi \times \pi',\psi) q^{-(n - 1) c(\pi) (s - 1/2)}$</span> for some nonnegative integer <span class="math-container">$c(\pi)$</span>, and it is precisely this nonnegative integer <span class="math-container">$c(\pi)$</span> that satisfies <span class="math-container">$\pi^{K_1(\mathfrak{p}^{c(\pi)})} \ni W^{\circ}$</span> and <span class="math-container">$\pi^{K_1(\mathfrak{p}^m)} = \{0\}$</span> for all <span class="math-container">$m &lt; c(\pi)$</span>.</p> <p>I don't know if the fact that (2) implies (1) has appeared anywhere in print, but I do know how to prove it. There is a <a href="https://projecteuclid.org/euclid.jmsj/1390600834" rel="nofollow noreferrer">paper</a> by Miyauchi that shows that the (Whittaker) newform <span class="math-container">$W^{\circ}$</span> given by (2) is such that <span class="math-container">$$\Psi(s,W^{\circ}) = \int_{F^{\times}} W^{\circ} \begin{pmatrix} x &amp; 0 &amp; \cdots &amp; 0 \\ 0 &amp; 1 &amp; \cdots &amp; 0 \\ \vdots &amp; \vdots &amp; \ddots &amp; \vdots \\ 0 &amp; &amp; \cdots &amp; 1 \end{pmatrix} |x|^{s - \frac{n - 1}{2}} \, d^{\times} x$$</span> is equal to the <span class="math-container">$L$</span>-function <span class="math-container">$L(s,\pi)$</span>, and the same method of proof (using Hecke operators) shows that <span class="math-container">$\Psi(s,W^{\circ},W'^{\circ}) = L(s, \pi \times \pi')$</span> for all spherical representations <span class="math-container">$\pi'$</span> of <span class="math-container">$\mathrm{GL}_{n-1}(F)$</span>. Now you can work backwards to show that <span class="math-container">$\epsilon(s,\pi \times \pi',\psi)$</span> is equal to <span class="math-container">$\epsilon(1/2, \pi \times \pi',\psi) q^{-(n - 1) c(\pi) (s - 1/2)}$</span>, as in Martin Dickson's answer.</p> Answer 4: <p>Yes this is true, but I don't think it's completely obvious. The following argument is taken from Roberts--Schmidt ``Local newforms for $GSp_4$'', which uses the same argument as the $GL_2$ case.</p> <p>Jacquet--Langlands proves the local functional equation $$ Z(1-s, \pi(w_0) W) = \gamma(s, W) Z(s, W) $$ where $W$ is in the Whittaker model, $Z(s, W)$ is the usual zeta integral, and $w_0 = \begin{bmatrix} &amp; -1 \\ 1 \end{bmatrix}$ is the long Weyl element. For any positive integer $n$, define $w_n = \begin{bmatrix} &amp; -1 \\ \varpi^n \end{bmatrix}$; we can write the above as $$ Z(1-s, \pi(w_n) W) = q^{n(s-1/2)} \gamma(s, W) Z(s, W). $$</p> <p>Now take $\pi$ of conductor $\mathfrak{p}^n$ (in the depth sense), and $W$ a generator for the space of $K(\mathfrak{p}^n)$-fixed vectors. It is known (presumably implicit in J--L) that one can choose $W$ s.t. $Z(s, W) = L(s, \pi)$. Also, one easily sees that $\pi(w_n)$ normalises $K(\mathfrak{p}^n)$ and hence acts on its fixed vectors, and so the LHS is also the corresponding local $L$-function, times the eigenvalue of $w_n$. Now dividing by the local $L$-functions to get the definition of the $\epsilon$-factor, one now sees that the conductor in the $\epsilon$-factor is what it should be.</p>
Closed formula for sine powers
260,326
7
https://mathoverflow.net/questions/260326/closed-formula-for-sine-powers
<p>I am looking for a closed formula for the expressions $$ \sum_{k=1}^{n-1} \sin\left(\pi \frac{k}{n}\right)^m,$$ with $n \in \mathbb{N}$ and $m \in \mathbb{N}$ odd.</p> <p>Playing with these sums a bit, I got the impression that the result is always a fairly simple algebraic expression, the dependence on $m$, $n$ however looking pretty arbitrary. </p>
[ { "answer_id": 260328, "score": 16, "text": "<p>You are looking at the sum of $m$-th powers of the imaginary parts of $n$-th roots of $-1.$ First, note that these can be expressed as polynomials in elementary symmetric functions (of the imaginary parts), so you just need to find the polynomial of which they are roots. Writing \n$$(x+i y)^n +1 = 0,$$ you get two equations, in $x$ and $y,$ eliminate $x$ and you will get the polynomial you want.</p>\n" }, { "answer_id": 260373, "score": 5, "text": "<p>Letting $\\xi=e^{\\frac{2\\pi i}{2n}}$, a primitive root of unity $z^{2n}=1$. Then, $\\sin\\left(\\frac{\\pi k}n\\right)=\\frac{\\xi^{k}-\\xi^{-k}}{2i}$ and hence\n\\begin{align}\n\\sum_{k=0}^{n-1}\\sin^m\\left(\\frac{\\pi k}n\\right)\n&amp;=\\frac1{(2i)^m}\\sum_{k=0}^{n-1}\\sum_{j=0}^m\\binom{m}j(-1)^j\\xi^{-jk}\\xi^{(m-j)k} \\\\\n&amp;=\\frac1{(2i)^m}\\sum_{j=0}^m(-1)^j\\binom{m}j\\sum_{k=0}^{n-1}\\xi^{(m-2j)k} \\\\\n&amp;=\\frac1{(2i)^m}\\sum_{j=0}^m(-1)^j\\binom{m}j\\frac{\\xi^{(m-2j)n}-1}{\\xi^{m-2j}-1} \\\\\n&amp;=\\frac1{(2i)^m}\\sum_{j=0}^m(-1)^j\\binom{m}j\\frac{\\xi^{mn}-1}{\\xi^{m-2j}-1}.\n\\end{align}\nThe outcome is <em>parity-dependent</em>. For example, if $m\\rightarrow 2m$ is even then the average value is\n$$\\frac1n\\sum_{k=0}^{n-1}\\sin^{2m}\\left(\\frac{\\pi k}n\\right)\n=\\frac1{2^{2m}}\\binom{2m}m,$$\nas long as $n&gt;m$. This is because $\\xi^{2mn}-1=0$ and the only time $\\xi^{2m-2j}-1=0$ is when $j=m$ (since $n&gt;m$). Therefore,\n$$\\sum_{k=0}^{n-1}\\sin^{2m}\\left(\\frac{\\pi k}n\\right)\n=\\frac1{(2i)^{2m}}\\cdot (-1)^m\\binom{2m}m=\\frac1{2^{2m}}\\binom{2m}m.$$\nBy the way, \n$$\\frac1{\\pi}\\int_0^{\\pi}\\sin^{2m}x\\,dx=\\frac1{2^{2m}}\\binom{2m}m.$$\nContrary to what I thought initially, things are a bit tricky (or unlikely) that there is a closed formula in the case $m\\rightarrow 2m+1$ is odd. What we get with the above approach is just an identity:\n$$\\sum_{k=0}^{n-1}\\sin^{2m+1}\\left(\\frac{\\pi k}n\\right)\n=\\frac1{2^{2m}}\\sum_{j=0}^m(-1)^j\\binom{2m+1}{m-j}\\cot\\left(\\frac{2j+1}{2n}\\pi\\right).$$ </p>\n" }, { "answer_id": 260328, "score": 16, "text": "<p>You are looking at the sum of $m$-th powers of the imaginary parts of $n$-th roots of $-1.$ First, note that these can be expressed as polynomials in elementary symmetric functions (of the imaginary parts), so you just need to find the polynomial of which they are roots. Writing \n$$(x+i y)^n +1 = 0,$$ you get two equations, in $x$ and $y,$ eliminate $x$ and you will get the polynomial you want.</p>\n" }, { "answer_id": 260373, "score": 5, "text": "<p>Letting $\\xi=e^{\\frac{2\\pi i}{2n}}$, a primitive root of unity $z^{2n}=1$. Then, $\\sin\\left(\\frac{\\pi k}n\\right)=\\frac{\\xi^{k}-\\xi^{-k}}{2i}$ and hence\n\\begin{align}\n\\sum_{k=0}^{n-1}\\sin^m\\left(\\frac{\\pi k}n\\right)\n&amp;=\\frac1{(2i)^m}\\sum_{k=0}^{n-1}\\sum_{j=0}^m\\binom{m}j(-1)^j\\xi^{-jk}\\xi^{(m-j)k} \\\\\n&amp;=\\frac1{(2i)^m}\\sum_{j=0}^m(-1)^j\\binom{m}j\\sum_{k=0}^{n-1}\\xi^{(m-2j)k} \\\\\n&amp;=\\frac1{(2i)^m}\\sum_{j=0}^m(-1)^j\\binom{m}j\\frac{\\xi^{(m-2j)n}-1}{\\xi^{m-2j}-1} \\\\\n&amp;=\\frac1{(2i)^m}\\sum_{j=0}^m(-1)^j\\binom{m}j\\frac{\\xi^{mn}-1}{\\xi^{m-2j}-1}.\n\\end{align}\nThe outcome is <em>parity-dependent</em>. For example, if $m\\rightarrow 2m$ is even then the average value is\n$$\\frac1n\\sum_{k=0}^{n-1}\\sin^{2m}\\left(\\frac{\\pi k}n\\right)\n=\\frac1{2^{2m}}\\binom{2m}m,$$\nas long as $n&gt;m$. This is because $\\xi^{2mn}-1=0$ and the only time $\\xi^{2m-2j}-1=0$ is when $j=m$ (since $n&gt;m$). Therefore,\n$$\\sum_{k=0}^{n-1}\\sin^{2m}\\left(\\frac{\\pi k}n\\right)\n=\\frac1{(2i)^{2m}}\\cdot (-1)^m\\binom{2m}m=\\frac1{2^{2m}}\\binom{2m}m.$$\nBy the way, \n$$\\frac1{\\pi}\\int_0^{\\pi}\\sin^{2m}x\\,dx=\\frac1{2^{2m}}\\binom{2m}m.$$\nContrary to what I thought initially, things are a bit tricky (or unlikely) that there is a closed formula in the case $m\\rightarrow 2m+1$ is odd. What we get with the above approach is just an identity:\n$$\\sum_{k=0}^{n-1}\\sin^{2m+1}\\left(\\frac{\\pi k}n\\right)\n=\\frac1{2^{2m}}\\sum_{j=0}^m(-1)^j\\binom{2m+1}{m-j}\\cot\\left(\\frac{2j+1}{2n}\\pi\\right).$$ </p>\n" } ]
Question: Closed formula for sine powers Answer 1: <p>You are looking at the sum of $m$-th powers of the imaginary parts of $n$-th roots of $-1.$ First, note that these can be expressed as polynomials in elementary symmetric functions (of the imaginary parts), so you just need to find the polynomial of which they are roots. Writing $$(x+i y)^n +1 = 0,$$ you get two equations, in $x$ and $y,$ eliminate $x$ and you will get the polynomial you want.</p> Answer 2: <p>Letting $\xi=e^{\frac{2\pi i}{2n}}$, a primitive root of unity $z^{2n}=1$. Then, $\sin\left(\frac{\pi k}n\right)=\frac{\xi^{k}-\xi^{-k}}{2i}$ and hence \begin{align} \sum_{k=0}^{n-1}\sin^m\left(\frac{\pi k}n\right) &amp;=\frac1{(2i)^m}\sum_{k=0}^{n-1}\sum_{j=0}^m\binom{m}j(-1)^j\xi^{-jk}\xi^{(m-j)k} \\ &amp;=\frac1{(2i)^m}\sum_{j=0}^m(-1)^j\binom{m}j\sum_{k=0}^{n-1}\xi^{(m-2j)k} \\ &amp;=\frac1{(2i)^m}\sum_{j=0}^m(-1)^j\binom{m}j\frac{\xi^{(m-2j)n}-1}{\xi^{m-2j}-1} \\ &amp;=\frac1{(2i)^m}\sum_{j=0}^m(-1)^j\binom{m}j\frac{\xi^{mn}-1}{\xi^{m-2j}-1}. \end{align} The outcome is <em>parity-dependent</em>. For example, if $m\rightarrow 2m$ is even then the average value is $$\frac1n\sum_{k=0}^{n-1}\sin^{2m}\left(\frac{\pi k}n\right) =\frac1{2^{2m}}\binom{2m}m,$$ as long as $n&gt;m$. This is because $\xi^{2mn}-1=0$ and the only time $\xi^{2m-2j}-1=0$ is when $j=m$ (since $n&gt;m$). Therefore, $$\sum_{k=0}^{n-1}\sin^{2m}\left(\frac{\pi k}n\right) =\frac1{(2i)^{2m}}\cdot (-1)^m\binom{2m}m=\frac1{2^{2m}}\binom{2m}m.$$ By the way, $$\frac1{\pi}\int_0^{\pi}\sin^{2m}x\,dx=\frac1{2^{2m}}\binom{2m}m.$$ Contrary to what I thought initially, things are a bit tricky (or unlikely) that there is a closed formula in the case $m\rightarrow 2m+1$ is odd. What we get with the above approach is just an identity: $$\sum_{k=0}^{n-1}\sin^{2m+1}\left(\frac{\pi k}n\right) =\frac1{2^{2m}}\sum_{j=0}^m(-1)^j\binom{2m+1}{m-j}\cot\left(\frac{2j+1}{2n}\pi\right).$$ </p> Answer 3: <p>You are looking at the sum of $m$-th powers of the imaginary parts of $n$-th roots of $-1.$ First, note that these can be expressed as polynomials in elementary symmetric functions (of the imaginary parts), so you just need to find the polynomial of which they are roots. Writing $$(x+i y)^n +1 = 0,$$ you get two equations, in $x$ and $y,$ eliminate $x$ and you will get the polynomial you want.</p> Answer 4: <p>Letting $\xi=e^{\frac{2\pi i}{2n}}$, a primitive root of unity $z^{2n}=1$. Then, $\sin\left(\frac{\pi k}n\right)=\frac{\xi^{k}-\xi^{-k}}{2i}$ and hence \begin{align} \sum_{k=0}^{n-1}\sin^m\left(\frac{\pi k}n\right) &amp;=\frac1{(2i)^m}\sum_{k=0}^{n-1}\sum_{j=0}^m\binom{m}j(-1)^j\xi^{-jk}\xi^{(m-j)k} \\ &amp;=\frac1{(2i)^m}\sum_{j=0}^m(-1)^j\binom{m}j\sum_{k=0}^{n-1}\xi^{(m-2j)k} \\ &amp;=\frac1{(2i)^m}\sum_{j=0}^m(-1)^j\binom{m}j\frac{\xi^{(m-2j)n}-1}{\xi^{m-2j}-1} \\ &amp;=\frac1{(2i)^m}\sum_{j=0}^m(-1)^j\binom{m}j\frac{\xi^{mn}-1}{\xi^{m-2j}-1}. \end{align} The outcome is <em>parity-dependent</em>. For example, if $m\rightarrow 2m$ is even then the average value is $$\frac1n\sum_{k=0}^{n-1}\sin^{2m}\left(\frac{\pi k}n\right) =\frac1{2^{2m}}\binom{2m}m,$$ as long as $n&gt;m$. This is because $\xi^{2mn}-1=0$ and the only time $\xi^{2m-2j}-1=0$ is when $j=m$ (since $n&gt;m$). Therefore, $$\sum_{k=0}^{n-1}\sin^{2m}\left(\frac{\pi k}n\right) =\frac1{(2i)^{2m}}\cdot (-1)^m\binom{2m}m=\frac1{2^{2m}}\binom{2m}m.$$ By the way, $$\frac1{\pi}\int_0^{\pi}\sin^{2m}x\,dx=\frac1{2^{2m}}\binom{2m}m.$$ Contrary to what I thought initially, things are a bit tricky (or unlikely) that there is a closed formula in the case $m\rightarrow 2m+1$ is odd. What we get with the above approach is just an identity: $$\sum_{k=0}^{n-1}\sin^{2m+1}\left(\frac{\pi k}n\right) =\frac1{2^{2m}}\sum_{j=0}^m(-1)^j\binom{2m+1}{m-j}\cot\left(\frac{2j+1}{2n}\pi\right).$$ </p>
Are there recursive sets $X$ with Property A that contain infinitely many incompressible strings?
258,843
6
https://mathoverflow.net/questions/258843/are-there-recursive-sets-x-with-property-a-that-contain-infinitely-many-incomp
<p>Let us say a set $X$ satisfies Property A if$$\liminf_{n \to \infty} {{\left|X^{\le n}\right|}\over n} = 0.$$Are there recursive sets $X$ satisfying Property A that contain infinitely many incompressible strings?</p>
[ { "answer_id": 258850, "score": 5, "text": "<p>I assume that $X$ is a set of binary strings and the binary strings are coded by natural numbers in this way: consider this enumeration of all strings: </p>\n\n<p>$\\lambda$, 0,1,00,01,10,11,000,.....</p>\n\n<p>(in which, at first we have the string with length zero, then all strings of length one in alphabetical orde and so on).Now code the $n$ th member of this enumeration by the number $n$.</p>\n\n<p>Now we define the set $X$ as $\\{w~:~ |w|=2^i~ for ~some ~i\\in \\mathbb{N}\\}$ ($|w|$ denotes the length of the string $w$).</p>\n\n<p>It can easily checked that $X$ satisfies the property $\\textsf{A}$. It is also well-known that we have incompressible strings of any length (because the number \nof Turing machines of length&lt;$n$ is at most $2^n-1$ but we have $2^n$ string of length $n$). Therefore $X$ contains infinitely many incompressible strings. </p>\n" }, { "answer_id": 258860, "score": 3, "text": "<p>I'm assuming you call `incompressible' integers whose binary representation $x$ satisfy $K(x)&gt;|x|-c$ for a fixed $c$? </p>\n\n<p>If that's the case, the answer is no. If $X$ is a computable set of integers and some integer $k$ of representation $x$, with $|x|=N$, belongs to $X$, then $K(x|N) \\leq \\log |X^{\\leq 2^{N+1}}| + c'$ for a fixed $c'$. (Indeed one can describe $x$ by just giving its position as a string of length $N$ inside $X$). But by your assumption on $X$, $\\log |X^{\\leq 2^{N+1}}| - N$ tends to $-\\infty$, so almost all strings in $X$ satisfy, say, $K(x\\mid|x|) &lt; |x| - 3c$. You can then remove the the condition $|x|$ and get $K(x) &lt; |x| - 2c$ by a usual padding technique (see the book by Li and Vitanyi, I can elaborate if this is important). </p>\n" }, { "answer_id": 258850, "score": 5, "text": "<p>I assume that $X$ is a set of binary strings and the binary strings are coded by natural numbers in this way: consider this enumeration of all strings: </p>\n\n<p>$\\lambda$, 0,1,00,01,10,11,000,.....</p>\n\n<p>(in which, at first we have the string with length zero, then all strings of length one in alphabetical orde and so on).Now code the $n$ th member of this enumeration by the number $n$.</p>\n\n<p>Now we define the set $X$ as $\\{w~:~ |w|=2^i~ for ~some ~i\\in \\mathbb{N}\\}$ ($|w|$ denotes the length of the string $w$).</p>\n\n<p>It can easily checked that $X$ satisfies the property $\\textsf{A}$. It is also well-known that we have incompressible strings of any length (because the number \nof Turing machines of length&lt;$n$ is at most $2^n-1$ but we have $2^n$ string of length $n$). Therefore $X$ contains infinitely many incompressible strings. </p>\n" }, { "answer_id": 258860, "score": 3, "text": "<p>I'm assuming you call `incompressible' integers whose binary representation $x$ satisfy $K(x)&gt;|x|-c$ for a fixed $c$? </p>\n\n<p>If that's the case, the answer is no. If $X$ is a computable set of integers and some integer $k$ of representation $x$, with $|x|=N$, belongs to $X$, then $K(x|N) \\leq \\log |X^{\\leq 2^{N+1}}| + c'$ for a fixed $c'$. (Indeed one can describe $x$ by just giving its position as a string of length $N$ inside $X$). But by your assumption on $X$, $\\log |X^{\\leq 2^{N+1}}| - N$ tends to $-\\infty$, so almost all strings in $X$ satisfy, say, $K(x\\mid|x|) &lt; |x| - 3c$. You can then remove the the condition $|x|$ and get $K(x) &lt; |x| - 2c$ by a usual padding technique (see the book by Li and Vitanyi, I can elaborate if this is important). </p>\n" } ]
Question: Are there recursive sets $X$ with Property A that contain infinitely many incompressible strings? Answer 1: <p>I assume that $X$ is a set of binary strings and the binary strings are coded by natural numbers in this way: consider this enumeration of all strings: </p> <p>$\lambda$, 0,1,00,01,10,11,000,.....</p> <p>(in which, at first we have the string with length zero, then all strings of length one in alphabetical orde and so on).Now code the $n$ th member of this enumeration by the number $n$.</p> <p>Now we define the set $X$ as $\{w~:~ |w|=2^i~ for ~some ~i\in \mathbb{N}\}$ ($|w|$ denotes the length of the string $w$).</p> <p>It can easily checked that $X$ satisfies the property $\textsf{A}$. It is also well-known that we have incompressible strings of any length (because the number of Turing machines of length&lt;$n$ is at most $2^n-1$ but we have $2^n$ string of length $n$). Therefore $X$ contains infinitely many incompressible strings. </p> Answer 2: <p>I'm assuming you call `incompressible' integers whose binary representation $x$ satisfy $K(x)&gt;|x|-c$ for a fixed $c$? </p> <p>If that's the case, the answer is no. If $X$ is a computable set of integers and some integer $k$ of representation $x$, with $|x|=N$, belongs to $X$, then $K(x|N) \leq \log |X^{\leq 2^{N+1}}| + c'$ for a fixed $c'$. (Indeed one can describe $x$ by just giving its position as a string of length $N$ inside $X$). But by your assumption on $X$, $\log |X^{\leq 2^{N+1}}| - N$ tends to $-\infty$, so almost all strings in $X$ satisfy, say, $K(x\mid|x|) &lt; |x| - 3c$. You can then remove the the condition $|x|$ and get $K(x) &lt; |x| - 2c$ by a usual padding technique (see the book by Li and Vitanyi, I can elaborate if this is important). </p> Answer 3: <p>I assume that $X$ is a set of binary strings and the binary strings are coded by natural numbers in this way: consider this enumeration of all strings: </p> <p>$\lambda$, 0,1,00,01,10,11,000,.....</p> <p>(in which, at first we have the string with length zero, then all strings of length one in alphabetical orde and so on).Now code the $n$ th member of this enumeration by the number $n$.</p> <p>Now we define the set $X$ as $\{w~:~ |w|=2^i~ for ~some ~i\in \mathbb{N}\}$ ($|w|$ denotes the length of the string $w$).</p> <p>It can easily checked that $X$ satisfies the property $\textsf{A}$. It is also well-known that we have incompressible strings of any length (because the number of Turing machines of length&lt;$n$ is at most $2^n-1$ but we have $2^n$ string of length $n$). Therefore $X$ contains infinitely many incompressible strings. </p> Answer 4: <p>I'm assuming you call `incompressible' integers whose binary representation $x$ satisfy $K(x)&gt;|x|-c$ for a fixed $c$? </p> <p>If that's the case, the answer is no. If $X$ is a computable set of integers and some integer $k$ of representation $x$, with $|x|=N$, belongs to $X$, then $K(x|N) \leq \log |X^{\leq 2^{N+1}}| + c'$ for a fixed $c'$. (Indeed one can describe $x$ by just giving its position as a string of length $N$ inside $X$). But by your assumption on $X$, $\log |X^{\leq 2^{N+1}}| - N$ tends to $-\infty$, so almost all strings in $X$ satisfy, say, $K(x\mid|x|) &lt; |x| - 3c$. You can then remove the the condition $|x|$ and get $K(x) &lt; |x| - 2c$ by a usual padding technique (see the book by Li and Vitanyi, I can elaborate if this is important). </p>
Lecture notes, videos and other learning materials about $\infty$-category theory
258,859
6
https://mathoverflow.net/questions/258859/lecture-notes-videos-and-other-learning-materials-about-infty-category-theor
<p>I've heard several times (and realized myself) that Lurie's tomes (extraordinary as they are) are not so ideal for self study.</p> <p>I think it's a good idea to have some kind of compiled list of learning material about infinity categories (could be articles as well if they are considered to have a pedagogical style). Preferably they would have emphasis on intuition, and application rather than formality and rigor.</p> <p>Lecture notes, Videos and any other relevant form of studying material would help a lot in travelling this technical terrain. </p>
[ { "answer_id": 258870, "score": 7, "text": "<p>I found <a href=\"https://arxiv.org/abs/1303.4669\" rel=\"nofollow noreferrer\">A Whirlwind Tour of the World of (∞,1)-categories</a> by Omar Antolín Camarena (a student of Jacob Lurie) to be quite insightful for quasicategories.</p>\n" }, { "answer_id": 380883, "score": 5, "text": "<p>Here are some extra references. (Maybe this is easier to read: <a href=\"https://pastebin.com/raw/se05pF8E\" rel=\"noreferrer\">link</a>)</p>\n<pre><code>├── Complete Segal Spaces\n│   ├── [BC, 10 Pages] Equivariant Complete Segal Spaces.pdf\n│   ├── [Brito, 26 Pages] Segal Objects and the Grothendieck Construction.pdf\n│   ├── [Osborne, 25 Pages] Decomposing the Classifying Diagram in Terms of Classifying Spaces of Groups.pdf\n│   ├── [Rasekh, 17 Pages] A Model for the Higher Category of Higher Categories.pdf\n│   ├── [Rasekh, 44 Pages] Complete Segal Objects.pdf\n│   ├── [Rasekh, 478 Pages] A Theory of Elementary Higher Toposes.pdf\n│   ├── [Rasekh, 61 Pages] Introduction to Complete Segal Spaces.pdf\n│   ├── [Rasekh, 75 Pages] Yoneda Lemma for Simplicial Spaces.pdf\n│   ├── [Rasekh, 81 Pages] Cartesian Fibrations and Representability.pdf\n│   ├── [Stenzel, 21 Pages] Univalence and Completeness of Segal Objects.pdf\n│   └── [Stenzel, 26 Pages] Bousfield-Segal Spaces.pdf\n├── Elementary ∞-Topoi\n│   ├── [Rasekh, 10 Pages] Yoneda Lemma for Elementary Higher Toposes.pdf\n│   ├── [Rasekh, 30 Pages] Filter Quotients and Non-Presentable (∞,1)-Toposes.pdf\n│   ├── [Rasekh, 40 Pages] A Theory of Elementary Higher Toposes.pdf\n│   ├── [Rasekh, 51 Pages] Every Elementary Higher Topos Has a Natural Number Object.pdf\n│   └── [Rasekh, 84 Pages] An Elementary Approach to Truncations.pdf\n├── Enriched (∞,1)-Categories\n│   ├── [AMR, 68 Pages] Factorization Homology of Enriched ∞-Categories.pdf\n│   ├── [GH, 100 Pages] Enriched ∞-Categories via Non-Symmetric ∞-Operads.pdf\n│   ├── [Haugseng, 29 Pages] Bimodules and Natural Transformations for Enriched ∞-Categories.pdf\n│   └── [Haugseng, 52 Pages] Rectification of Enriched Infinity-Categories.pdf\n├── General References\n│   ├── [AL, 26 Pages] Exponentiable Higher Toposes.pdf\n│   ├── [Arctaedius, 38 Pages] Grothendieck's Homotopy Hypothesis and the Homotopy Theory of Homotopy Theories.pdf\n│   ├── [Bergner, 13 Pages] A Survey of (∞, 1)-Categories.pdf\n│   ├── [Bergner, 287 Pages] The Homotopy Theory of (∞,1)-Categories.pdf\n│   ├── [Bergner, 29 Pages] A Survey of Models for (∞,n)-Categories.pdf\n│   ├── [Bergner, 39 Pages] Workshop on the Homotopy Theory of Homotopy Theories.pdf\n│   ├── [Camarena, 45 Pages] A Whirlwind Tour of the World of (∞,1)-Categories.pdf\n│   ├── [Clough, 35 Pages] An Outline of the Theory of (∞,1)-Categories.pdf\n│   ├── [Dorn, 99 Pages] Basic concepts in homotopy theory.pdf\n│   ├── [HF, 148 Pages] The Homotopy Theory of (∞,1)-Categories.pdf\n│   ├── [Mazel-Gee, 26 Pages] The Zen of ∞-Categories.pdf\n│   ├── [Porter, 37 Pages] 𝒮-categories, 𝒮-groupoids, Segal categories and quasicategories.pdf\n│   ├── [Porter, 54 Pages] Spaces as ∞-groupoids.pdf\n│   ├── [Porter, 759 Pages] The Crossed Menagerie.pdf\n│   ├── [Schommer-Pries, 65 Pages] Dualizability in Low-Dimensional Higher Category Theory.pdf\n│   └── [Simpson, 653 Pages] Homotopy Theory of Higher Categories.pdf\n├── Model (∞,1)-Categories\n│   ├── [LM, 21 Pages] From Fractions to Complete Segal Spaces.pdf\n│   ├── [Mazel-Gee, 16 Pages] A User's Guide to Co⧸Cartesian Fibrations.pdf\n│   ├── [Mazel-Gee, 20 Pages] Quillen Adjunctions Induce Adjunctions of Quasicategories.pdf\n│   ├── [Mazel-Gee, 26 Pages] The Universality of the Rezk Nerve.pdf\n│   ├── [Mazel-Gee, 29 Pages] Model ∞-Categories II: Quillen Adjunctions.pdf\n│   ├── [Mazel-Gee, 34 Pages] Model ∞-Categories III: The Fundamental Theorem.pdf\n│   ├── [Mazel-Gee, 41 Pages] All About the Grothendieck Construction.pdf\n│   ├── [Mazel-Gee, 43 Pages] Hammocks and Fractions in Relative ∞-Categories.pdf\n│   ├── [Mazel-Gee, 528 Slides] Goerss–Hopkins obstruction theory for ∞-Categories.pdf\n│   ├── [Mazel-Gee, 545 Pages] Goerss–Hopkins obstruction theory via model ∞-categories.pdf\n│   ├── [Mazel-Gee, 54 Pages] Goerss–Hopkins Obstruction Theory for ∞-Categories.pdf\n│   ├── [Mazel-Gee, 66 Pages] Model ∞-Categories I: Some Pleasant Properties of the ∞-Category of Simplicial Spaces.pdf\n│   └── [Mazel-Gee, 6 Pages] 𝔼_∞ Automorphisms of Motivic Morava E-Theories.pdf\n├── Other Models\n│   ├── Batanin ∞-Categories\n│   │   └── [Ara, 168 Pages] Sur les ∞-groupoïdes de Grothendieck et une variante ∞-catégorique.pdf\n│   ├── Grothendieck–Maltsiniotis ∞-Categories\n│   │   ├── [AL, 65 Pages] The Folk Model Category Structure on Strict ω-Categories Is Monoidal.pdf\n│   │   ├── [AM, 123 Pages] A Quillen's Theorem A for Strict ∞-Categories II: The ∞-Categorical Proof.pdf\n│   │   ├── [AM, 227 Pages] Join and Slices for Strict ∞-Categories.pdf\n│   │   ├── [AM, 25 Pages] The Brown–Golasinski Model Structure on Strict ∞-Groupoids Revisited.pdf\n│   │   ├── [AM, 42 Pages] Comparison of the n-Categorical Nerves.pdf\n│   │   ├── [AM, 51 Pages] A Quillen's Theorem A for Strict ∞-Categories I: The Simplicial Proof.pdf\n│   │   ├── [AM, 68 Pages] The Homotopy Type of the ∞-Category Associated to a Simplicial Complex.pdf\n│   │   ├── [AM, 92 Pages] Towards a Thomason Model Structure on the Category of Strict n-Categories.pdf\n│   │   ├── [Ara, 22 Pages] On Homotopy Types Modelized by Strict ∞-Groupoids.pdf\n│   │   ├── [Ara, 22 Pages] Strict ∞-Groupoids Are Grothendieck ∞-Groupoids.pdf\n│   │   ├── [Ara, 33 Pages] A Quillen Theorem B for Strict ∞-Categories.pdf\n│   │   ├── [Ara, 41 Pages] The Groupoidal Analogue Θ to Joyal's Category Θ Is a Test Category.pdf\n│   │   └── [Ara, 58 Pages] On the Homotopy Theory of Grothendieck ∞-Groupoids.pdf\n│   ├── Miscellany\n│   │   ├── [CL, 61 Pages] Weak ∞-Categories via Terminal Coalgebras.pdf\n│   │   ├── [Harpaz, 69 Pages] Quasi-Unital ∞-Categories.pdf\n│   │   ├── [LM, 68 Pages] Linear Quasi-Categories as Templicial Modules.pdf\n│   │   └── [Nikolaus, 26 Pages] Algebraic Models for Higher Categories.pdf\n│   ├── Relative Categories\n│   │   ├── [BK, 19 Pages] n-Relative Categories.pdf\n│   │   ├── [BK, 27 Pages] Relative Categories: Another Model for the Homotopy Theory of Homotopy Theories.pdf\n│   │   ├── [BK, 3 Pages] A Thomason-Like Quillen Equivalence Between Quasi-Categories and Relative Categories.pdf\n│   │   ├── [BK, 5 Pages] In the Category of Relative Categories the Rezk Equivalences Are Exactly the DK-equivalences.pdf\n│   │   └── [Meier, 21 Pages] Fibration Categories Are Fibrant Relative Categories.pdf\n│   └── Topological Categories\n│   ├── [Amrani, 22 Pages] A Model Structure on the Category of Topological Categories.pdf\n├── Parametrised (∞,1)-Categories\n│   ├── [BDGNS, 11 Pages] Parametrized Higher Category Theory and Higher Algebra: A General Introduction.pdf\n│   ├── [BDGNS, 23 Pages] Parametrized Higher Category Theory and Higher Algebra: Exposé I -- Elements of Parametrized Higher Category Theory.pdf\n│   ├── [Nardin, 21 Pages] Parametrized Higher Category Theory and Higher Algebra: Exposé IV -- Stability With Respect to an Orbital ∞-Category.pdf\n│   ├── [Shah, 81 Pages] Parametrized Higher Category Theory and Higher Algebra: Exposé II - Indexed Homotopy Limits and Colimits.pdf\n│   └── [Shah, 86 Slides] Parametrized Higher Category Theory.pdf\n├── Quasicategories\n│   ├── [AL, 26 Pages] Exponentiable Higher Toposes.pdf\n│   ├── [Beardsley, 78 Pages] Coalgebraic Structure and Intermediate Hopf–Galois Extensions of Thom Spectra in Quasicategories.pdf\n│   ├── [Berman, 13 Pages] On Lax Limits in ∞-Categories.pdf\n│   ├── [BG, 11 Pages] On the Fibrewise Effective Burnside ∞-Category.pdf\n│   ├── [BM, 30 Pages] Spectral Sequences in (∞,1)-Categories.pdf\n│   ├── [BS, 18 Pages] Fibrations in ∞-Category Theory.pdf\n│   ├── [BV, 267 Pages] Homotopy Invariant Algebraic Structures On Topological Spaces.djvu\n│   ├── [Campbell, 3 Pages] A Counterexample in Quasi-Category Theory.pdf\n│   ├── [Cisinski, 204 Pages] Algèbre Homotopique et Catégories Supérieures.pdf\n│   ├── [Cisinski, 446 Pages] Higher Categories and Homotopical Algebra.pdf\n│   ├── [Cisisnki, 57 Pages] Catégories Supérieures et Théorie des Topos.pdf\n│   ├── [DS, 46 Pages] Mapping Spaces in Quasi-Categories.pdf\n│   ├── [Fiore, 24 Pages] Quasicategorical Adjunctions.pdf\n│   ├── [GR, 11 Pages] Simplified HTT 4.3.2.15.pdf\n│   ├── [GR, 85 Pages] Some Higher Algebra.pdf\n│   ├── [Gregoric, 21 Pages] Gregoric Blitzkrieg.pdf\n│   ├── [Groth, 77 Pages] A Short Course on ∞-Categories.pdf\n│   ├── [Harpaz, 116 Pages] Little Cube Algebras and Factorisation Homology.pdf\n│   ├── [Harpaz, 7 Pages] Limits, colimits and adjunctions in ∞-categories.pdf\n│   ├── [Haugseng, 12 Pages] On (Co)ends in ∞-Categories.pdf\n│   ├── [Haugseng, 28 Pages] A Fibrational Mate Correspondence for ∞-Categories.pdf\n│   ├── [Haugseng, 87 Pages] Introduction to ∞-Categories.pdf\n│   ├── [HH, 684 Pages] Higher Categories I &amp; II\n│   │   ├── 0 [Hebestreit, 13 Pages] A Fairytale.pdf\n│   │   ├── 10 [Hebestreit, 11 Pages] Fat and Thin Slices.pdf\n│   │   ├── 11 [Hebestreit, 43 Pages] Cartesian Fibrations.pdf\n│   │   ├── 12 [Hebestreit, 121 Pages] Straightening and Unstraightening.pdf\n│   │   ├── 12' [HH, 53 Pages] Straightening and Unstraightening (Heuts's Notes).pdf\n│   │   ├── 13 [Hebestreit, 14 Pages] Homotopy Colimits.pdf\n│   │   ├── 14 [Hebestreit, 10 Pages] Simplicial Model Categories.pdf\n│   │   ├── 15 [Hebestreit, 35 Pages] Yoneda's Lemma, Adjunctions and (Co)Limits.pdf\n│   │   ├── 1 [Hebestreit, 37 Pages] Categories.pdf\n│   │   ├── 2 [Hebestreit, 25 Pages] Simplicial Sets.pdf\n│   │   ├── 3 [Hebestreit, 10 Pages] Quasicategories.pdf\n│   │   ├── 4 [Hebestreit, 53 Pages] Simplicial Categories.pdf\n│   │   ├── 5 [Hebestreit, 38 Pages] Simplicial Homotopy Theory.pdf\n│   │   ├── 6 [Hebestreit, 57 Pages] Quasicategories and Anima.pdf\n│   │   ├── 7 [Hebestreit, 67 Pages] Equivalences, Equivalences, Equivalences.pdf\n│   │   ├── 8 [Hebestreit, 9 Pages] A Fairytale.pdf\n│   │   └── 9 [Hebestreit, 88 Pages] Localisations and Model Categories.pdf\n│   ├── [Hinich, 111 Pages] Lectures on ∞-Categories.pdf\n│   ├── [Joyal, 244 Pages] Notes on Quasi-Categories.pdf\n│   ├── [Joyal, 350 Pages] The Theory of Quasi-Categories and its Applications II.pdf\n│   ├── [Joyal, 479 Pages] The Theory of Quasi-Categories.pdf\n│   ├── [Lejay, 139 Pages] Algèbres à Factorisation et Topos Supérieurs Exponentiables.pdf\n│   ├── [Lurie, 60 Pages] On ∞-Topoi.pdf\n│   ├── [Lurie, 841 Pages] Kerodon.pdf\n│   ├── [Lurie, 949 Pages] Higher Topos Theory.pdf\n│   ├── [Lysenko, 208 Pages] Lysenko's comments to Gaitsgory–Rozenblyum.pdf\n│   ├── [Morel, 118 Pages] ∞-Categories.pdf\n│   ├── [Nguyen, 94 Pages] Theorems in Higher Category Theory and Applications.pdf\n│   ├── [NRS, 21 Pages] Adjoint Functor Theorems for ∞-Categories.pdf\n│   ├── [Porta, 100 Pages] Derived formal moduli problems.pdf\n│   ├── [Rezk, 11 Pages] Degenerate Edges of Cartesian Fibrations are Cartesian Edges.pdf\n│   ├── [Rezk, 175 Pages] Stuff About Quasicategories.pdf\n│   ├── [Rezk, 50 Pages] Toposes and homotopy toposes.pdf\n│   ├── [Riehl, 20 Pages] Quasi-Categories as (∞,1)-Categories.pdf\n│   ├── [Riehl, 292 Pages] Categorical homotopy theory.pdf\n│   ├── [Riehl, 9 Pages] Associativity Data in an (∞,1)-Category.pdf\n│   ├── [Rovelli, 38 Pages] Weighted Limits in an (∞,1)-Category.pdf\n│   ├── [RS, 151 Pages] Notes on Higher Categories.pdf\n│   ├── [RV, 33 Pages] Completeness Results for Quasi-Categories of Algebras, Homotopy Limits, and Related General Constructions.pdf\n│   ├── [Stevenson, 12 Pages] Stability for Inner Fibrations Revisited.pdf\n│   ├── [Tanaka, 14 Pages] Functors (Between ∞-Categories) That Aren't Strictly Unital.pdf\n│   ├── [Thanh, 46 Pages] Quasicategories.pdf\n│   ├── [Wong, 83 Pages] The Grothendieck construction in enriched, internal and ∞-Category Theory.pdf\n│   ├── [Zsámboki, 31 Pages] A summary of higher topos theory.pdf\n│   └── [無, 5 Pages] HTT ToC.pdf\n├── Simplicial Categories\n│   ├── [Bergner, 16 Pages] A model category structure on the category of simplicial categories.pdf\n│   ├── [Bergner, 22 Pages] Complete Segal Spaces Arising From Simplicial Categories.pdf\n│   ├── [Bergner, 40 Pages] Three models for the homotopy theory of homotopy theories.pdf\n│   ├── [Cordier, 21 Pages] Sur la notion de diagramme homotopiquement cohérent.pdf\n│   ├── [CP, 26 Pages] Vogt's Theorem on Categories of Homotopy Coherent Diagrams.pdf\n│   ├── [CP, 54 Pages] Homotopy Coherent Category Theory.pdf\n│   ├── [DS, 29 Pages] Rigidification of Quasi-Categories.pdf\n│   ├── [DS, 64 Pages] Mapping Spaces in Quasi-Categories.pdf\n│   ├── [Hinich, 23 Pages] Homotopy Coherent Nerve in Deformation Theory.pdf\n│   ├── [HK, 14 Pages] Mapping Spaces in Homotopy Coherent Nerves.pdf\n│   ├── [Joyal, 66 Pages] Quasi-Categories vs Simplicial Categories.pdf\n│   ├── [Riehl, 16 Pages] On the Structure of Simplicial Categories Associated to Quasi-Categories.pdf\n│   ├── [Riehl, 26 Pages] Homotopy Coherent Structures.pdf\n│   ├── [Riehl, 292 Pages] Categorical Homotopy Theory.pdf\n│   └── [Riehl, 7 Pages] Understanding the Homotopy Coherent Nerve.pdf\n├── Unicity\n│   ├── [Bergner, 16 Pages] Equivalence of Models for Equivariant (∞,1)-Categories.pdf\n│   ├── [BS, 47 Pages] On the Unicity of the Homotopy Theory of Higher Categories.pdf\n│   ├── [JT, 49 Pages] Quasi-Categories vs Segal Spaces.pdf\n│   ├── [Riehl, 58 Pages] Seminar Notes on the Barwick–Schommer–Pries Unicity Theorem.pdf\n│   └── [Toën, 32 Pages] Vers une axiomatisation de la théorie des catégories supérieures.pdf\n└── Un⧸Straightening\n ├── [AF, 89 Pages] Fibrations of ∞-Categories.pdf\n ├── [BGN, 19 Pages] Dualizing Cartesian and Cocartesian Fibrations.pdf\n ├── [BS, 18 Pages] Fibrations in ∞-Category Theory.pdf\n ├── [Campbell, 31 Slides] A modular proof of the straightening theorem.pdf\n ├── [GHN, 42 Pages] Lax Colimits and Free Fibrations in ∞-Categories.pdf\n ├── [GR, 85 Pages] Some Higher Algebra.pdf\n ├── [Hebestreit, 121 Pages] Straightening and Unstraightening.pdf\n ├── [HH, 53 Pages] Straightening and Unstraightening (Heuts's Notes).pdf\n ├── [HM, 17 Pages] Left Fibrations and Homotopy Colimits II.pdf\n ├── [HM, 27 Pages] Left Fibrations and Homotopy Colimits.pdf\n ├── [Mazel-Gee, 16 Pages] A User's Guide to (Co)Cartesian Fibrations.pdf\n ├── [Mazel-Gee, 41 Pages] All About the Grothendieck Construction.pdf\n ├── [Noel, 1 Page] Cartesian Model Structure.pdf\n ├── [PK, 43 Pages] Straightening and Unstraightening.pdf\n ├── [Richardson, 14 Pages] Mapping Spaces and Straightening-Unstraightening.pdf\n ├── [Ruit, 71 Pages] Grothendieck constructions in higher category theory.pdf\n ├── [Stevenson, 41 Pages] Model Structures for Correspondences and Bifibrations.pdf\n ├── [Stevenson, 49 Pages] Covariant Model Structures and Simplicial Localization.pdf\n └── [Wong, 83 Pages] The Grothendieck Constructionin Enriched, Internal and ∞-Category Theory.pdf\n</code></pre>\n" }, { "answer_id": 258872, "score": 3, "text": "<p>Moritz Groth put up some excellent lecture notes: <a href=\"https://arxiv.org/abs/1007.2925v2\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1007.2925v2</a></p>\n<p>If you are more categorically minded, Emily Riehl's book has a lot about quasi-categories:\n<a href=\"http://www.math.jhu.edu/%7Eeriehl/cathtpy.pdf\" rel=\"nofollow noreferrer\">http://www.math.jhu.edu/~eriehl/cathtpy.pdf</a>\nand her website has a lot of exposition about cosmoi, an approach to <span class=\"math-container\">$\\infty$</span>-categories via 2-category theory, and not requiring one to work in the setting of quasi-categories:\n<a href=\"http://www.math.jhu.edu/%7Eeriehl/\" rel=\"nofollow noreferrer\">http://www.math.jhu.edu/~eriehl/</a></p>\n" }, { "answer_id": 258897, "score": 3, "text": "<p>It depends on from what direction you are coming and in what direction you are wanting to go! </p>\n\n<p>As one of the developers of the basic theory from way back, I like to approach things via homotopy coherence as that is where the quasi-category approach comes from, and it also more or less guarantees some nice diagrams and situations that are slightly more geometric or visual. Have a look at my n-Lab page: \n<a href=\"https://ncatlab.org/timporter/show/simplicial+foundations+for+homotopy+coherence\" rel=\"nofollow noreferrer\">https://ncatlab.org/timporter/show/simplicial+foundations+for+homotopy+coherence</a></p>\n\n<p>Other material that includes intuitive approaches include my book with Heiner Kamps: Abstract Homotopy and Simple Homotopy Theory, World Scientific, 462pp (ISBN 981-02-1602-5)</p>\n\n<p>and the Cubo notes: Abstract Homotopy Theory, the interaction of category theory and homotopy theory, (survey article, updated version of lecture notes from summer school course at Bressanone), Cubo, 5 (2003)115-165, 2003)(and here: <a href=\"https://ncatlab.org/nlab/files/Abstract-Homotopy.pdf\" rel=\"nofollow noreferrer\">https://ncatlab.org/nlab/files/Abstract-Homotopy.pdf</a>).</p>\n\n<p>Those do NOT go very far but concentrate on the intuition and basic structure. They may be too elementary for you but are intended to be readable. You can also access a version of the crossed menagerie: <a href=\"https://ncatlab.org/timporter/show/crossed+menagerie\" rel=\"nofollow noreferrer\">https://ncatlab.org/timporter/show/crossed+menagerie</a>\nwhich discusses a lot of the relationship with non-abelian cohomology, if that is what is of interest to you. </p>\n\n<p>I hope this helps.</p>\n" }, { "answer_id": 432428, "score": 1, "text": "<p>There are now more resources since the question was raised. Markus Land's <a href=\"https://link.springer.com/book/10.1007/978-3-030-61524-6\" rel=\"nofollow noreferrer\"><em>Introduction to infinity-categories</em></a> grew out from lecture notes, and seemingly much better fitted for pedagogy.</p>\n<p>On the other hands, due (or thanks?) to pandemic, there are more video resources. For example, there is a series of lectures by Achim Krause and Thomas Nikolaus <a href=\"https://www.youtube.com/playlist?list=PLsmqTkj4MGTDenpj574aSvIRBROwCugoB\" rel=\"nofollow noreferrer\">recorded and uploaded on the YouTube channel</a>.</p>\n" }, { "answer_id": 258870, "score": 7, "text": "<p>I found <a href=\"https://arxiv.org/abs/1303.4669\" rel=\"nofollow noreferrer\">A Whirlwind Tour of the World of (∞,1)-categories</a> by Omar Antolín Camarena (a student of Jacob Lurie) to be quite insightful for quasicategories.</p>\n" }, { "answer_id": 380883, "score": 5, "text": "<p>Here are some extra references. (Maybe this is easier to read: <a href=\"https://pastebin.com/raw/se05pF8E\" rel=\"noreferrer\">link</a>)</p>\n<pre><code>├── Complete Segal Spaces\n│   ├── [BC, 10 Pages] Equivariant Complete Segal Spaces.pdf\n│   ├── [Brito, 26 Pages] Segal Objects and the Grothendieck Construction.pdf\n│   ├── [Osborne, 25 Pages] Decomposing the Classifying Diagram in Terms of Classifying Spaces of Groups.pdf\n│   ├── [Rasekh, 17 Pages] A Model for the Higher Category of Higher Categories.pdf\n│   ├── [Rasekh, 44 Pages] Complete Segal Objects.pdf\n│   ├── [Rasekh, 478 Pages] A Theory of Elementary Higher Toposes.pdf\n│   ├── [Rasekh, 61 Pages] Introduction to Complete Segal Spaces.pdf\n│   ├── [Rasekh, 75 Pages] Yoneda Lemma for Simplicial Spaces.pdf\n│   ├── [Rasekh, 81 Pages] Cartesian Fibrations and Representability.pdf\n│   ├── [Stenzel, 21 Pages] Univalence and Completeness of Segal Objects.pdf\n│   └── [Stenzel, 26 Pages] Bousfield-Segal Spaces.pdf\n├── Elementary ∞-Topoi\n│   ├── [Rasekh, 10 Pages] Yoneda Lemma for Elementary Higher Toposes.pdf\n│   ├── [Rasekh, 30 Pages] Filter Quotients and Non-Presentable (∞,1)-Toposes.pdf\n│   ├── [Rasekh, 40 Pages] A Theory of Elementary Higher Toposes.pdf\n│   ├── [Rasekh, 51 Pages] Every Elementary Higher Topos Has a Natural Number Object.pdf\n│   └── [Rasekh, 84 Pages] An Elementary Approach to Truncations.pdf\n├── Enriched (∞,1)-Categories\n│   ├── [AMR, 68 Pages] Factorization Homology of Enriched ∞-Categories.pdf\n│   ├── [GH, 100 Pages] Enriched ∞-Categories via Non-Symmetric ∞-Operads.pdf\n│   ├── [Haugseng, 29 Pages] Bimodules and Natural Transformations for Enriched ∞-Categories.pdf\n│   └── [Haugseng, 52 Pages] Rectification of Enriched Infinity-Categories.pdf\n├── General References\n│   ├── [AL, 26 Pages] Exponentiable Higher Toposes.pdf\n│   ├── [Arctaedius, 38 Pages] Grothendieck's Homotopy Hypothesis and the Homotopy Theory of Homotopy Theories.pdf\n│   ├── [Bergner, 13 Pages] A Survey of (∞, 1)-Categories.pdf\n│   ├── [Bergner, 287 Pages] The Homotopy Theory of (∞,1)-Categories.pdf\n│   ├── [Bergner, 29 Pages] A Survey of Models for (∞,n)-Categories.pdf\n│   ├── [Bergner, 39 Pages] Workshop on the Homotopy Theory of Homotopy Theories.pdf\n│   ├── [Camarena, 45 Pages] A Whirlwind Tour of the World of (∞,1)-Categories.pdf\n│   ├── [Clough, 35 Pages] An Outline of the Theory of (∞,1)-Categories.pdf\n│   ├── [Dorn, 99 Pages] Basic concepts in homotopy theory.pdf\n│   ├── [HF, 148 Pages] The Homotopy Theory of (∞,1)-Categories.pdf\n│   ├── [Mazel-Gee, 26 Pages] The Zen of ∞-Categories.pdf\n│   ├── [Porter, 37 Pages] 𝒮-categories, 𝒮-groupoids, Segal categories and quasicategories.pdf\n│   ├── [Porter, 54 Pages] Spaces as ∞-groupoids.pdf\n│   ├── [Porter, 759 Pages] The Crossed Menagerie.pdf\n│   ├── [Schommer-Pries, 65 Pages] Dualizability in Low-Dimensional Higher Category Theory.pdf\n│   └── [Simpson, 653 Pages] Homotopy Theory of Higher Categories.pdf\n├── Model (∞,1)-Categories\n│   ├── [LM, 21 Pages] From Fractions to Complete Segal Spaces.pdf\n│   ├── [Mazel-Gee, 16 Pages] A User's Guide to Co⧸Cartesian Fibrations.pdf\n│   ├── [Mazel-Gee, 20 Pages] Quillen Adjunctions Induce Adjunctions of Quasicategories.pdf\n│   ├── [Mazel-Gee, 26 Pages] The Universality of the Rezk Nerve.pdf\n│   ├── [Mazel-Gee, 29 Pages] Model ∞-Categories II: Quillen Adjunctions.pdf\n│   ├── [Mazel-Gee, 34 Pages] Model ∞-Categories III: The Fundamental Theorem.pdf\n│   ├── [Mazel-Gee, 41 Pages] All About the Grothendieck Construction.pdf\n│   ├── [Mazel-Gee, 43 Pages] Hammocks and Fractions in Relative ∞-Categories.pdf\n│   ├── [Mazel-Gee, 528 Slides] Goerss–Hopkins obstruction theory for ∞-Categories.pdf\n│   ├── [Mazel-Gee, 545 Pages] Goerss–Hopkins obstruction theory via model ∞-categories.pdf\n│   ├── [Mazel-Gee, 54 Pages] Goerss–Hopkins Obstruction Theory for ∞-Categories.pdf\n│   ├── [Mazel-Gee, 66 Pages] Model ∞-Categories I: Some Pleasant Properties of the ∞-Category of Simplicial Spaces.pdf\n│   └── [Mazel-Gee, 6 Pages] 𝔼_∞ Automorphisms of Motivic Morava E-Theories.pdf\n├── Other Models\n│   ├── Batanin ∞-Categories\n│   │   └── [Ara, 168 Pages] Sur les ∞-groupoïdes de Grothendieck et une variante ∞-catégorique.pdf\n│   ├── Grothendieck–Maltsiniotis ∞-Categories\n│   │   ├── [AL, 65 Pages] The Folk Model Category Structure on Strict ω-Categories Is Monoidal.pdf\n│   │   ├── [AM, 123 Pages] A Quillen's Theorem A for Strict ∞-Categories II: The ∞-Categorical Proof.pdf\n│   │   ├── [AM, 227 Pages] Join and Slices for Strict ∞-Categories.pdf\n│   │   ├── [AM, 25 Pages] The Brown–Golasinski Model Structure on Strict ∞-Groupoids Revisited.pdf\n│   │   ├── [AM, 42 Pages] Comparison of the n-Categorical Nerves.pdf\n│   │   ├── [AM, 51 Pages] A Quillen's Theorem A for Strict ∞-Categories I: The Simplicial Proof.pdf\n│   │   ├── [AM, 68 Pages] The Homotopy Type of the ∞-Category Associated to a Simplicial Complex.pdf\n│   │   ├── [AM, 92 Pages] Towards a Thomason Model Structure on the Category of Strict n-Categories.pdf\n│   │   ├── [Ara, 22 Pages] On Homotopy Types Modelized by Strict ∞-Groupoids.pdf\n│   │   ├── [Ara, 22 Pages] Strict ∞-Groupoids Are Grothendieck ∞-Groupoids.pdf\n│   │   ├── [Ara, 33 Pages] A Quillen Theorem B for Strict ∞-Categories.pdf\n│   │   ├── [Ara, 41 Pages] The Groupoidal Analogue Θ to Joyal's Category Θ Is a Test Category.pdf\n│   │   └── [Ara, 58 Pages] On the Homotopy Theory of Grothendieck ∞-Groupoids.pdf\n│   ├── Miscellany\n│   │   ├── [CL, 61 Pages] Weak ∞-Categories via Terminal Coalgebras.pdf\n│   │   ├── [Harpaz, 69 Pages] Quasi-Unital ∞-Categories.pdf\n│   │   ├── [LM, 68 Pages] Linear Quasi-Categories as Templicial Modules.pdf\n│   │   └── [Nikolaus, 26 Pages] Algebraic Models for Higher Categories.pdf\n│   ├── Relative Categories\n│   │   ├── [BK, 19 Pages] n-Relative Categories.pdf\n│   │   ├── [BK, 27 Pages] Relative Categories: Another Model for the Homotopy Theory of Homotopy Theories.pdf\n│   │   ├── [BK, 3 Pages] A Thomason-Like Quillen Equivalence Between Quasi-Categories and Relative Categories.pdf\n│   │   ├── [BK, 5 Pages] In the Category of Relative Categories the Rezk Equivalences Are Exactly the DK-equivalences.pdf\n│   │   └── [Meier, 21 Pages] Fibration Categories Are Fibrant Relative Categories.pdf\n│   └── Topological Categories\n│   ├── [Amrani, 22 Pages] A Model Structure on the Category of Topological Categories.pdf\n├── Parametrised (∞,1)-Categories\n│   ├── [BDGNS, 11 Pages] Parametrized Higher Category Theory and Higher Algebra: A General Introduction.pdf\n│   ├── [BDGNS, 23 Pages] Parametrized Higher Category Theory and Higher Algebra: Exposé I -- Elements of Parametrized Higher Category Theory.pdf\n│   ├── [Nardin, 21 Pages] Parametrized Higher Category Theory and Higher Algebra: Exposé IV -- Stability With Respect to an Orbital ∞-Category.pdf\n│   ├── [Shah, 81 Pages] Parametrized Higher Category Theory and Higher Algebra: Exposé II - Indexed Homotopy Limits and Colimits.pdf\n│   └── [Shah, 86 Slides] Parametrized Higher Category Theory.pdf\n├── Quasicategories\n│   ├── [AL, 26 Pages] Exponentiable Higher Toposes.pdf\n│   ├── [Beardsley, 78 Pages] Coalgebraic Structure and Intermediate Hopf–Galois Extensions of Thom Spectra in Quasicategories.pdf\n│   ├── [Berman, 13 Pages] On Lax Limits in ∞-Categories.pdf\n│   ├── [BG, 11 Pages] On the Fibrewise Effective Burnside ∞-Category.pdf\n│   ├── [BM, 30 Pages] Spectral Sequences in (∞,1)-Categories.pdf\n│   ├── [BS, 18 Pages] Fibrations in ∞-Category Theory.pdf\n│   ├── [BV, 267 Pages] Homotopy Invariant Algebraic Structures On Topological Spaces.djvu\n│   ├── [Campbell, 3 Pages] A Counterexample in Quasi-Category Theory.pdf\n│   ├── [Cisinski, 204 Pages] Algèbre Homotopique et Catégories Supérieures.pdf\n│   ├── [Cisinski, 446 Pages] Higher Categories and Homotopical Algebra.pdf\n│   ├── [Cisisnki, 57 Pages] Catégories Supérieures et Théorie des Topos.pdf\n│   ├── [DS, 46 Pages] Mapping Spaces in Quasi-Categories.pdf\n│   ├── [Fiore, 24 Pages] Quasicategorical Adjunctions.pdf\n│   ├── [GR, 11 Pages] Simplified HTT 4.3.2.15.pdf\n│   ├── [GR, 85 Pages] Some Higher Algebra.pdf\n│   ├── [Gregoric, 21 Pages] Gregoric Blitzkrieg.pdf\n│   ├── [Groth, 77 Pages] A Short Course on ∞-Categories.pdf\n│   ├── [Harpaz, 116 Pages] Little Cube Algebras and Factorisation Homology.pdf\n│   ├── [Harpaz, 7 Pages] Limits, colimits and adjunctions in ∞-categories.pdf\n│   ├── [Haugseng, 12 Pages] On (Co)ends in ∞-Categories.pdf\n│   ├── [Haugseng, 28 Pages] A Fibrational Mate Correspondence for ∞-Categories.pdf\n│   ├── [Haugseng, 87 Pages] Introduction to ∞-Categories.pdf\n│   ├── [HH, 684 Pages] Higher Categories I &amp; II\n│   │   ├── 0 [Hebestreit, 13 Pages] A Fairytale.pdf\n│   │   ├── 10 [Hebestreit, 11 Pages] Fat and Thin Slices.pdf\n│   │   ├── 11 [Hebestreit, 43 Pages] Cartesian Fibrations.pdf\n│   │   ├── 12 [Hebestreit, 121 Pages] Straightening and Unstraightening.pdf\n│   │   ├── 12' [HH, 53 Pages] Straightening and Unstraightening (Heuts's Notes).pdf\n│   │   ├── 13 [Hebestreit, 14 Pages] Homotopy Colimits.pdf\n│   │   ├── 14 [Hebestreit, 10 Pages] Simplicial Model Categories.pdf\n│   │   ├── 15 [Hebestreit, 35 Pages] Yoneda's Lemma, Adjunctions and (Co)Limits.pdf\n│   │   ├── 1 [Hebestreit, 37 Pages] Categories.pdf\n│   │   ├── 2 [Hebestreit, 25 Pages] Simplicial Sets.pdf\n│   │   ├── 3 [Hebestreit, 10 Pages] Quasicategories.pdf\n│   │   ├── 4 [Hebestreit, 53 Pages] Simplicial Categories.pdf\n│   │   ├── 5 [Hebestreit, 38 Pages] Simplicial Homotopy Theory.pdf\n│   │   ├── 6 [Hebestreit, 57 Pages] Quasicategories and Anima.pdf\n│   │   ├── 7 [Hebestreit, 67 Pages] Equivalences, Equivalences, Equivalences.pdf\n│   │   ├── 8 [Hebestreit, 9 Pages] A Fairytale.pdf\n│   │   └── 9 [Hebestreit, 88 Pages] Localisations and Model Categories.pdf\n│   ├── [Hinich, 111 Pages] Lectures on ∞-Categories.pdf\n│   ├── [Joyal, 244 Pages] Notes on Quasi-Categories.pdf\n│   ├── [Joyal, 350 Pages] The Theory of Quasi-Categories and its Applications II.pdf\n│   ├── [Joyal, 479 Pages] The Theory of Quasi-Categories.pdf\n│   ├── [Lejay, 139 Pages] Algèbres à Factorisation et Topos Supérieurs Exponentiables.pdf\n│   ├── [Lurie, 60 Pages] On ∞-Topoi.pdf\n│   ├── [Lurie, 841 Pages] Kerodon.pdf\n│   ├── [Lurie, 949 Pages] Higher Topos Theory.pdf\n│   ├── [Lysenko, 208 Pages] Lysenko's comments to Gaitsgory–Rozenblyum.pdf\n│   ├── [Morel, 118 Pages] ∞-Categories.pdf\n│   ├── [Nguyen, 94 Pages] Theorems in Higher Category Theory and Applications.pdf\n│   ├── [NRS, 21 Pages] Adjoint Functor Theorems for ∞-Categories.pdf\n│   ├── [Porta, 100 Pages] Derived formal moduli problems.pdf\n│   ├── [Rezk, 11 Pages] Degenerate Edges of Cartesian Fibrations are Cartesian Edges.pdf\n│   ├── [Rezk, 175 Pages] Stuff About Quasicategories.pdf\n│   ├── [Rezk, 50 Pages] Toposes and homotopy toposes.pdf\n│   ├── [Riehl, 20 Pages] Quasi-Categories as (∞,1)-Categories.pdf\n│   ├── [Riehl, 292 Pages] Categorical homotopy theory.pdf\n│   ├── [Riehl, 9 Pages] Associativity Data in an (∞,1)-Category.pdf\n│   ├── [Rovelli, 38 Pages] Weighted Limits in an (∞,1)-Category.pdf\n│   ├── [RS, 151 Pages] Notes on Higher Categories.pdf\n│   ├── [RV, 33 Pages] Completeness Results for Quasi-Categories of Algebras, Homotopy Limits, and Related General Constructions.pdf\n│   ├── [Stevenson, 12 Pages] Stability for Inner Fibrations Revisited.pdf\n│   ├── [Tanaka, 14 Pages] Functors (Between ∞-Categories) That Aren't Strictly Unital.pdf\n│   ├── [Thanh, 46 Pages] Quasicategories.pdf\n│   ├── [Wong, 83 Pages] The Grothendieck construction in enriched, internal and ∞-Category Theory.pdf\n│   ├── [Zsámboki, 31 Pages] A summary of higher topos theory.pdf\n│   └── [無, 5 Pages] HTT ToC.pdf\n├── Simplicial Categories\n│   ├── [Bergner, 16 Pages] A model category structure on the category of simplicial categories.pdf\n│   ├── [Bergner, 22 Pages] Complete Segal Spaces Arising From Simplicial Categories.pdf\n│   ├── [Bergner, 40 Pages] Three models for the homotopy theory of homotopy theories.pdf\n│   ├── [Cordier, 21 Pages] Sur la notion de diagramme homotopiquement cohérent.pdf\n│   ├── [CP, 26 Pages] Vogt's Theorem on Categories of Homotopy Coherent Diagrams.pdf\n│   ├── [CP, 54 Pages] Homotopy Coherent Category Theory.pdf\n│   ├── [DS, 29 Pages] Rigidification of Quasi-Categories.pdf\n│   ├── [DS, 64 Pages] Mapping Spaces in Quasi-Categories.pdf\n│   ├── [Hinich, 23 Pages] Homotopy Coherent Nerve in Deformation Theory.pdf\n│   ├── [HK, 14 Pages] Mapping Spaces in Homotopy Coherent Nerves.pdf\n│   ├── [Joyal, 66 Pages] Quasi-Categories vs Simplicial Categories.pdf\n│   ├── [Riehl, 16 Pages] On the Structure of Simplicial Categories Associated to Quasi-Categories.pdf\n│   ├── [Riehl, 26 Pages] Homotopy Coherent Structures.pdf\n│   ├── [Riehl, 292 Pages] Categorical Homotopy Theory.pdf\n│   └── [Riehl, 7 Pages] Understanding the Homotopy Coherent Nerve.pdf\n├── Unicity\n│   ├── [Bergner, 16 Pages] Equivalence of Models for Equivariant (∞,1)-Categories.pdf\n│   ├── [BS, 47 Pages] On the Unicity of the Homotopy Theory of Higher Categories.pdf\n│   ├── [JT, 49 Pages] Quasi-Categories vs Segal Spaces.pdf\n│   ├── [Riehl, 58 Pages] Seminar Notes on the Barwick–Schommer–Pries Unicity Theorem.pdf\n│   └── [Toën, 32 Pages] Vers une axiomatisation de la théorie des catégories supérieures.pdf\n└── Un⧸Straightening\n ├── [AF, 89 Pages] Fibrations of ∞-Categories.pdf\n ├── [BGN, 19 Pages] Dualizing Cartesian and Cocartesian Fibrations.pdf\n ├── [BS, 18 Pages] Fibrations in ∞-Category Theory.pdf\n ├── [Campbell, 31 Slides] A modular proof of the straightening theorem.pdf\n ├── [GHN, 42 Pages] Lax Colimits and Free Fibrations in ∞-Categories.pdf\n ├── [GR, 85 Pages] Some Higher Algebra.pdf\n ├── [Hebestreit, 121 Pages] Straightening and Unstraightening.pdf\n ├── [HH, 53 Pages] Straightening and Unstraightening (Heuts's Notes).pdf\n ├── [HM, 17 Pages] Left Fibrations and Homotopy Colimits II.pdf\n ├── [HM, 27 Pages] Left Fibrations and Homotopy Colimits.pdf\n ├── [Mazel-Gee, 16 Pages] A User's Guide to (Co)Cartesian Fibrations.pdf\n ├── [Mazel-Gee, 41 Pages] All About the Grothendieck Construction.pdf\n ├── [Noel, 1 Page] Cartesian Model Structure.pdf\n ├── [PK, 43 Pages] Straightening and Unstraightening.pdf\n ├── [Richardson, 14 Pages] Mapping Spaces and Straightening-Unstraightening.pdf\n ├── [Ruit, 71 Pages] Grothendieck constructions in higher category theory.pdf\n ├── [Stevenson, 41 Pages] Model Structures for Correspondences and Bifibrations.pdf\n ├── [Stevenson, 49 Pages] Covariant Model Structures and Simplicial Localization.pdf\n └── [Wong, 83 Pages] The Grothendieck Constructionin Enriched, Internal and ∞-Category Theory.pdf\n</code></pre>\n" }, { "answer_id": 258872, "score": 3, "text": "<p>Moritz Groth put up some excellent lecture notes: <a href=\"https://arxiv.org/abs/1007.2925v2\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1007.2925v2</a></p>\n<p>If you are more categorically minded, Emily Riehl's book has a lot about quasi-categories:\n<a href=\"http://www.math.jhu.edu/%7Eeriehl/cathtpy.pdf\" rel=\"nofollow noreferrer\">http://www.math.jhu.edu/~eriehl/cathtpy.pdf</a>\nand her website has a lot of exposition about cosmoi, an approach to <span class=\"math-container\">$\\infty$</span>-categories via 2-category theory, and not requiring one to work in the setting of quasi-categories:\n<a href=\"http://www.math.jhu.edu/%7Eeriehl/\" rel=\"nofollow noreferrer\">http://www.math.jhu.edu/~eriehl/</a></p>\n" }, { "answer_id": 258897, "score": 3, "text": "<p>It depends on from what direction you are coming and in what direction you are wanting to go! </p>\n\n<p>As one of the developers of the basic theory from way back, I like to approach things via homotopy coherence as that is where the quasi-category approach comes from, and it also more or less guarantees some nice diagrams and situations that are slightly more geometric or visual. Have a look at my n-Lab page: \n<a href=\"https://ncatlab.org/timporter/show/simplicial+foundations+for+homotopy+coherence\" rel=\"nofollow noreferrer\">https://ncatlab.org/timporter/show/simplicial+foundations+for+homotopy+coherence</a></p>\n\n<p>Other material that includes intuitive approaches include my book with Heiner Kamps: Abstract Homotopy and Simple Homotopy Theory, World Scientific, 462pp (ISBN 981-02-1602-5)</p>\n\n<p>and the Cubo notes: Abstract Homotopy Theory, the interaction of category theory and homotopy theory, (survey article, updated version of lecture notes from summer school course at Bressanone), Cubo, 5 (2003)115-165, 2003)(and here: <a href=\"https://ncatlab.org/nlab/files/Abstract-Homotopy.pdf\" rel=\"nofollow noreferrer\">https://ncatlab.org/nlab/files/Abstract-Homotopy.pdf</a>).</p>\n\n<p>Those do NOT go very far but concentrate on the intuition and basic structure. They may be too elementary for you but are intended to be readable. You can also access a version of the crossed menagerie: <a href=\"https://ncatlab.org/timporter/show/crossed+menagerie\" rel=\"nofollow noreferrer\">https://ncatlab.org/timporter/show/crossed+menagerie</a>\nwhich discusses a lot of the relationship with non-abelian cohomology, if that is what is of interest to you. </p>\n\n<p>I hope this helps.</p>\n" }, { "answer_id": 432428, "score": 1, "text": "<p>There are now more resources since the question was raised. Markus Land's <a href=\"https://link.springer.com/book/10.1007/978-3-030-61524-6\" rel=\"nofollow noreferrer\"><em>Introduction to infinity-categories</em></a> grew out from lecture notes, and seemingly much better fitted for pedagogy.</p>\n<p>On the other hands, due (or thanks?) to pandemic, there are more video resources. For example, there is a series of lectures by Achim Krause and Thomas Nikolaus <a href=\"https://www.youtube.com/playlist?list=PLsmqTkj4MGTDenpj574aSvIRBROwCugoB\" rel=\"nofollow noreferrer\">recorded and uploaded on the YouTube channel</a>.</p>\n" } ]
Question: Lecture notes, videos and other learning materials about $\infty$-category theory Answer 1: <p>I found <a href="https://arxiv.org/abs/1303.4669" rel="nofollow noreferrer">A Whirlwind Tour of the World of (∞,1)-categories</a> by Omar Antolín Camarena (a student of Jacob Lurie) to be quite insightful for quasicategories.</p> Answer 2: <p>Here are some extra references. (Maybe this is easier to read: <a href="https://pastebin.com/raw/se05pF8E" rel="noreferrer">link</a>)</p> <pre><code>├── Complete Segal Spaces │   ├── [BC, 10 Pages] Equivariant Complete Segal Spaces.pdf │   ├── [Brito, 26 Pages] Segal Objects and the Grothendieck Construction.pdf │   ├── [Osborne, 25 Pages] Decomposing the Classifying Diagram in Terms of Classifying Spaces of Groups.pdf │   ├── [Rasekh, 17 Pages] A Model for the Higher Category of Higher Categories.pdf │   ├── [Rasekh, 44 Pages] Complete Segal Objects.pdf │   ├── [Rasekh, 478 Pages] A Theory of Elementary Higher Toposes.pdf │   ├── [Rasekh, 61 Pages] Introduction to Complete Segal Spaces.pdf │   ├── [Rasekh, 75 Pages] Yoneda Lemma for Simplicial Spaces.pdf │   ├── [Rasekh, 81 Pages] Cartesian Fibrations and Representability.pdf │   ├── [Stenzel, 21 Pages] Univalence and Completeness of Segal Objects.pdf │   └── [Stenzel, 26 Pages] Bousfield-Segal Spaces.pdf ├── Elementary ∞-Topoi │   ├── [Rasekh, 10 Pages] Yoneda Lemma for Elementary Higher Toposes.pdf │   ├── [Rasekh, 30 Pages] Filter Quotients and Non-Presentable (∞,1)-Toposes.pdf │   ├── [Rasekh, 40 Pages] A Theory of Elementary Higher Toposes.pdf │   ├── [Rasekh, 51 Pages] Every Elementary Higher Topos Has a Natural Number Object.pdf │   └── [Rasekh, 84 Pages] An Elementary Approach to Truncations.pdf ├── Enriched (∞,1)-Categories │   ├── [AMR, 68 Pages] Factorization Homology of Enriched ∞-Categories.pdf │   ├── [GH, 100 Pages] Enriched ∞-Categories via Non-Symmetric ∞-Operads.pdf │   ├── [Haugseng, 29 Pages] Bimodules and Natural Transformations for Enriched ∞-Categories.pdf │   └── [Haugseng, 52 Pages] Rectification of Enriched Infinity-Categories.pdf ├── General References │   ├── [AL, 26 Pages] Exponentiable Higher Toposes.pdf │   ├── [Arctaedius, 38 Pages] Grothendieck's Homotopy Hypothesis and the Homotopy Theory of Homotopy Theories.pdf │   ├── [Bergner, 13 Pages] A Survey of (∞, 1)-Categories.pdf │   ├── [Bergner, 287 Pages] The Homotopy Theory of (∞,1)-Categories.pdf │   ├── [Bergner, 29 Pages] A Survey of Models for (∞,n)-Categories.pdf │   ├── [Bergner, 39 Pages] Workshop on the Homotopy Theory of Homotopy Theories.pdf │   ├── [Camarena, 45 Pages] A Whirlwind Tour of the World of (∞,1)-Categories.pdf │   ├── [Clough, 35 Pages] An Outline of the Theory of (∞,1)-Categories.pdf │   ├── [Dorn, 99 Pages] Basic concepts in homotopy theory.pdf │   ├── [HF, 148 Pages] The Homotopy Theory of (∞,1)-Categories.pdf │   ├── [Mazel-Gee, 26 Pages] The Zen of ∞-Categories.pdf │   ├── [Porter, 37 Pages] 𝒮-categories, 𝒮-groupoids, Segal categories and quasicategories.pdf │   ├── [Porter, 54 Pages] Spaces as ∞-groupoids.pdf │   ├── [Porter, 759 Pages] The Crossed Menagerie.pdf │   ├── [Schommer-Pries, 65 Pages] Dualizability in Low-Dimensional Higher Category Theory.pdf │   └── [Simpson, 653 Pages] Homotopy Theory of Higher Categories.pdf ├── Model (∞,1)-Categories │   ├── [LM, 21 Pages] From Fractions to Complete Segal Spaces.pdf │   ├── [Mazel-Gee, 16 Pages] A User's Guide to Co⧸Cartesian Fibrations.pdf │   ├── [Mazel-Gee, 20 Pages] Quillen Adjunctions Induce Adjunctions of Quasicategories.pdf │   ├── [Mazel-Gee, 26 Pages] The Universality of the Rezk Nerve.pdf │   ├── [Mazel-Gee, 29 Pages] Model ∞-Categories II: Quillen Adjunctions.pdf │   ├── [Mazel-Gee, 34 Pages] Model ∞-Categories III: The Fundamental Theorem.pdf │   ├── [Mazel-Gee, 41 Pages] All About the Grothendieck Construction.pdf │   ├── [Mazel-Gee, 43 Pages] Hammocks and Fractions in Relative ∞-Categories.pdf │   ├── [Mazel-Gee, 528 Slides] Goerss–Hopkins obstruction theory for ∞-Categories.pdf │   ├── [Mazel-Gee, 545 Pages] Goerss–Hopkins obstruction theory via model ∞-categories.pdf │   ├── [Mazel-Gee, 54 Pages] Goerss–Hopkins Obstruction Theory for ∞-Categories.pdf │   ├── [Mazel-Gee, 66 Pages] Model ∞-Categories I: Some Pleasant Properties of the ∞-Category of Simplicial Spaces.pdf │   └── [Mazel-Gee, 6 Pages] 𝔼_∞ Automorphisms of Motivic Morava E-Theories.pdf ├── Other Models │   ├── Batanin ∞-Categories │   │   └── [Ara, 168 Pages] Sur les ∞-groupoïdes de Grothendieck et une variante ∞-catégorique.pdf │   ├── Grothendieck–Maltsiniotis ∞-Categories │   │   ├── [AL, 65 Pages] The Folk Model Category Structure on Strict ω-Categories Is Monoidal.pdf │   │   ├── [AM, 123 Pages] A Quillen's Theorem A for Strict ∞-Categories II: The ∞-Categorical Proof.pdf │   │   ├── [AM, 227 Pages] Join and Slices for Strict ∞-Categories.pdf │   │   ├── [AM, 25 Pages] The Brown–Golasinski Model Structure on Strict ∞-Groupoids Revisited.pdf │   │   ├── [AM, 42 Pages] Comparison of the n-Categorical Nerves.pdf │   │   ├── [AM, 51 Pages] A Quillen's Theorem A for Strict ∞-Categories I: The Simplicial Proof.pdf │   │   ├── [AM, 68 Pages] The Homotopy Type of the ∞-Category Associated to a Simplicial Complex.pdf │   │   ├── [AM, 92 Pages] Towards a Thomason Model Structure on the Category of Strict n-Categories.pdf │   │   ├── [Ara, 22 Pages] On Homotopy Types Modelized by Strict ∞-Groupoids.pdf │   │   ├── [Ara, 22 Pages] Strict ∞-Groupoids Are Grothendieck ∞-Groupoids.pdf │   │   ├── [Ara, 33 Pages] A Quillen Theorem B for Strict ∞-Categories.pdf │   │   ├── [Ara, 41 Pages] The Groupoidal Analogue Θ to Joyal's Category Θ Is a Test Category.pdf │   │   └── [Ara, 58 Pages] On the Homotopy Theory of Grothendieck ∞-Groupoids.pdf │   ├── Miscellany │   │   ├── [CL, 61 Pages] Weak ∞-Categories via Terminal Coalgebras.pdf │   │   ├── [Harpaz, 69 Pages] Quasi-Unital ∞-Categories.pdf │   │   ├── [LM, 68 Pages] Linear Quasi-Categories as Templicial Modules.pdf │   │   └── [Nikolaus, 26 Pages] Algebraic Models for Higher Categories.pdf │   ├── Relative Categories │   │   ├── [BK, 19 Pages] n-Relative Categories.pdf │   │   ├── [BK, 27 Pages] Relative Categories: Another Model for the Homotopy Theory of Homotopy Theories.pdf │   │   ├── [BK, 3 Pages] A Thomason-Like Quillen Equivalence Between Quasi-Categories and Relative Categories.pdf │   │   ├── [BK, 5 Pages] In the Category of Relative Categories the Rezk Equivalences Are Exactly the DK-equivalences.pdf │   │   └── [Meier, 21 Pages] Fibration Categories Are Fibrant Relative Categories.pdf │   └── Topological Categories │   ├── [Amrani, 22 Pages] A Model Structure on the Category of Topological Categories.pdf ├── Parametrised (∞,1)-Categories │   ├── [BDGNS, 11 Pages] Parametrized Higher Category Theory and Higher Algebra: A General Introduction.pdf │   ├── [BDGNS, 23 Pages] Parametrized Higher Category Theory and Higher Algebra: Exposé I -- Elements of Parametrized Higher Category Theory.pdf │   ├── [Nardin, 21 Pages] Parametrized Higher Category Theory and Higher Algebra: Exposé IV -- Stability With Respect to an Orbital ∞-Category.pdf │   ├── [Shah, 81 Pages] Parametrized Higher Category Theory and Higher Algebra: Exposé II - Indexed Homotopy Limits and Colimits.pdf │   └── [Shah, 86 Slides] Parametrized Higher Category Theory.pdf ├── Quasicategories │   ├── [AL, 26 Pages] Exponentiable Higher Toposes.pdf │   ├── [Beardsley, 78 Pages] Coalgebraic Structure and Intermediate Hopf–Galois Extensions of Thom Spectra in Quasicategories.pdf │   ├── [Berman, 13 Pages] On Lax Limits in ∞-Categories.pdf │   ├── [BG, 11 Pages] On the Fibrewise Effective Burnside ∞-Category.pdf │   ├── [BM, 30 Pages] Spectral Sequences in (∞,1)-Categories.pdf │   ├── [BS, 18 Pages] Fibrations in ∞-Category Theory.pdf │   ├── [BV, 267 Pages] Homotopy Invariant Algebraic Structures On Topological Spaces.djvu │   ├── [Campbell, 3 Pages] A Counterexample in Quasi-Category Theory.pdf │   ├── [Cisinski, 204 Pages] Algèbre Homotopique et Catégories Supérieures.pdf │   ├── [Cisinski, 446 Pages] Higher Categories and Homotopical Algebra.pdf │   ├── [Cisisnki, 57 Pages] Catégories Supérieures et Théorie des Topos.pdf │   ├── [DS, 46 Pages] Mapping Spaces in Quasi-Categories.pdf │   ├── [Fiore, 24 Pages] Quasicategorical Adjunctions.pdf │   ├── [GR, 11 Pages] Simplified HTT 4.3.2.15.pdf │   ├── [GR, 85 Pages] Some Higher Algebra.pdf │   ├── [Gregoric, 21 Pages] Gregoric Blitzkrieg.pdf │   ├── [Groth, 77 Pages] A Short Course on ∞-Categories.pdf │   ├── [Harpaz, 116 Pages] Little Cube Algebras and Factorisation Homology.pdf │   ├── [Harpaz, 7 Pages] Limits, colimits and adjunctions in ∞-categories.pdf │   ├── [Haugseng, 12 Pages] On (Co)ends in ∞-Categories.pdf │   ├── [Haugseng, 28 Pages] A Fibrational Mate Correspondence for ∞-Categories.pdf │   ├── [Haugseng, 87 Pages] Introduction to ∞-Categories.pdf │   ├── [HH, 684 Pages] Higher Categories I &amp; II │   │   ├── 0 [Hebestreit, 13 Pages] A Fairytale.pdf │   │   ├── 10 [Hebestreit, 11 Pages] Fat and Thin Slices.pdf │   │   ├── 11 [Hebestreit, 43 Pages] Cartesian Fibrations.pdf │   │   ├── 12 [Hebestreit, 121 Pages] Straightening and Unstraightening.pdf │   │   ├── 12' [HH, 53 Pages] Straightening and Unstraightening (Heuts's Notes).pdf │   │   ├── 13 [Hebestreit, 14 Pages] Homotopy Colimits.pdf │   │   ├── 14 [Hebestreit, 10 Pages] Simplicial Model Categories.pdf │   │   ├── 15 [Hebestreit, 35 Pages] Yoneda's Lemma, Adjunctions and (Co)Limits.pdf │   │   ├── 1 [Hebestreit, 37 Pages] Categories.pdf │   │   ├── 2 [Hebestreit, 25 Pages] Simplicial Sets.pdf │   │   ├── 3 [Hebestreit, 10 Pages] Quasicategories.pdf │   │   ├── 4 [Hebestreit, 53 Pages] Simplicial Categories.pdf │   │   ├── 5 [Hebestreit, 38 Pages] Simplicial Homotopy Theory.pdf │   │   ├── 6 [Hebestreit, 57 Pages] Quasicategories and Anima.pdf │   │   ├── 7 [Hebestreit, 67 Pages] Equivalences, Equivalences, Equivalences.pdf │   │   ├── 8 [Hebestreit, 9 Pages] A Fairytale.pdf │   │   └── 9 [Hebestreit, 88 Pages] Localisations and Model Categories.pdf │   ├── [Hinich, 111 Pages] Lectures on ∞-Categories.pdf │   ├── [Joyal, 244 Pages] Notes on Quasi-Categories.pdf │   ├── [Joyal, 350 Pages] The Theory of Quasi-Categories and its Applications II.pdf │   ├── [Joyal, 479 Pages] The Theory of Quasi-Categories.pdf │   ├── [Lejay, 139 Pages] Algèbres à Factorisation et Topos Supérieurs Exponentiables.pdf │   ├── [Lurie, 60 Pages] On ∞-Topoi.pdf │   ├── [Lurie, 841 Pages] Kerodon.pdf │   ├── [Lurie, 949 Pages] Higher Topos Theory.pdf │   ├── [Lysenko, 208 Pages] Lysenko's comments to Gaitsgory–Rozenblyum.pdf │   ├── [Morel, 118 Pages] ∞-Categories.pdf │   ├── [Nguyen, 94 Pages] Theorems in Higher Category Theory and Applications.pdf │   ├── [NRS, 21 Pages] Adjoint Functor Theorems for ∞-Categories.pdf │   ├── [Porta, 100 Pages] Derived formal moduli problems.pdf │   ├── [Rezk, 11 Pages] Degenerate Edges of Cartesian Fibrations are Cartesian Edges.pdf │   ├── [Rezk, 175 Pages] Stuff About Quasicategories.pdf │   ├── [Rezk, 50 Pages] Toposes and homotopy toposes.pdf │   ├── [Riehl, 20 Pages] Quasi-Categories as (∞,1)-Categories.pdf │   ├── [Riehl, 292 Pages] Categorical homotopy theory.pdf │   ├── [Riehl, 9 Pages] Associativity Data in an (∞,1)-Category.pdf │   ├── [Rovelli, 38 Pages] Weighted Limits in an (∞,1)-Category.pdf │   ├── [RS, 151 Pages] Notes on Higher Categories.pdf │   ├── [RV, 33 Pages] Completeness Results for Quasi-Categories of Algebras, Homotopy Limits, and Related General Constructions.pdf │   ├── [Stevenson, 12 Pages] Stability for Inner Fibrations Revisited.pdf │   ├── [Tanaka, 14 Pages] Functors (Between ∞-Categories) That Aren't Strictly Unital.pdf │   ├── [Thanh, 46 Pages] Quasicategories.pdf │   ├── [Wong, 83 Pages] The Grothendieck construction in enriched, internal and ∞-Category Theory.pdf │   ├── [Zsámboki, 31 Pages] A summary of higher topos theory.pdf │   └── [無, 5 Pages] HTT ToC.pdf ├── Simplicial Categories │   ├── [Bergner, 16 Pages] A model category structure on the category of simplicial categories.pdf │   ├── [Bergner, 22 Pages] Complete Segal Spaces Arising From Simplicial Categories.pdf │   ├── [Bergner, 40 Pages] Three models for the homotopy theory of homotopy theories.pdf │   ├── [Cordier, 21 Pages] Sur la notion de diagramme homotopiquement cohérent.pdf │   ├── [CP, 26 Pages] Vogt's Theorem on Categories of Homotopy Coherent Diagrams.pdf │   ├── [CP, 54 Pages] Homotopy Coherent Category Theory.pdf │   ├── [DS, 29 Pages] Rigidification of Quasi-Categories.pdf │   ├── [DS, 64 Pages] Mapping Spaces in Quasi-Categories.pdf │   ├── [Hinich, 23 Pages] Homotopy Coherent Nerve in Deformation Theory.pdf │   ├── [HK, 14 Pages] Mapping Spaces in Homotopy Coherent Nerves.pdf │   ├── [Joyal, 66 Pages] Quasi-Categories vs Simplicial Categories.pdf │   ├── [Riehl, 16 Pages] On the Structure of Simplicial Categories Associated to Quasi-Categories.pdf │   ├── [Riehl, 26 Pages] Homotopy Coherent Structures.pdf │   ├── [Riehl, 292 Pages] Categorical Homotopy Theory.pdf │   └── [Riehl, 7 Pages] Understanding the Homotopy Coherent Nerve.pdf ├── Unicity │   ├── [Bergner, 16 Pages] Equivalence of Models for Equivariant (∞,1)-Categories.pdf │   ├── [BS, 47 Pages] On the Unicity of the Homotopy Theory of Higher Categories.pdf │   ├── [JT, 49 Pages] Quasi-Categories vs Segal Spaces.pdf │   ├── [Riehl, 58 Pages] Seminar Notes on the Barwick–Schommer–Pries Unicity Theorem.pdf │   └── [Toën, 32 Pages] Vers une axiomatisation de la théorie des catégories supérieures.pdf └── Un⧸Straightening ├── [AF, 89 Pages] Fibrations of ∞-Categories.pdf ├── [BGN, 19 Pages] Dualizing Cartesian and Cocartesian Fibrations.pdf ├── [BS, 18 Pages] Fibrations in ∞-Category Theory.pdf ├── [Campbell, 31 Slides] A modular proof of the straightening theorem.pdf ├── [GHN, 42 Pages] Lax Colimits and Free Fibrations in ∞-Categories.pdf ├── [GR, 85 Pages] Some Higher Algebra.pdf ├── [Hebestreit, 121 Pages] Straightening and Unstraightening.pdf ├── [HH, 53 Pages] Straightening and Unstraightening (Heuts's Notes).pdf ├── [HM, 17 Pages] Left Fibrations and Homotopy Colimits II.pdf ├── [HM, 27 Pages] Left Fibrations and Homotopy Colimits.pdf ├── [Mazel-Gee, 16 Pages] A User's Guide to (Co)Cartesian Fibrations.pdf ├── [Mazel-Gee, 41 Pages] All About the Grothendieck Construction.pdf ├── [Noel, 1 Page] Cartesian Model Structure.pdf ├── [PK, 43 Pages] Straightening and Unstraightening.pdf ├── [Richardson, 14 Pages] Mapping Spaces and Straightening-Unstraightening.pdf ├── [Ruit, 71 Pages] Grothendieck constructions in higher category theory.pdf ├── [Stevenson, 41 Pages] Model Structures for Correspondences and Bifibrations.pdf ├── [Stevenson, 49 Pages] Covariant Model Structures and Simplicial Localization.pdf └── [Wong, 83 Pages] The Grothendieck Constructionin Enriched, Internal and ∞-Category Theory.pdf </code></pre> Answer 3: <p>Moritz Groth put up some excellent lecture notes: <a href="https://arxiv.org/abs/1007.2925v2" rel="nofollow noreferrer">https://arxiv.org/abs/1007.2925v2</a></p> <p>If you are more categorically minded, Emily Riehl's book has a lot about quasi-categories: <a href="http://www.math.jhu.edu/%7Eeriehl/cathtpy.pdf" rel="nofollow noreferrer">http://www.math.jhu.edu/~eriehl/cathtpy.pdf</a> and her website has a lot of exposition about cosmoi, an approach to <span class="math-container">$\infty$</span>-categories via 2-category theory, and not requiring one to work in the setting of quasi-categories: <a href="http://www.math.jhu.edu/%7Eeriehl/" rel="nofollow noreferrer">http://www.math.jhu.edu/~eriehl/</a></p> Answer 4: <p>It depends on from what direction you are coming and in what direction you are wanting to go! </p> <p>As one of the developers of the basic theory from way back, I like to approach things via homotopy coherence as that is where the quasi-category approach comes from, and it also more or less guarantees some nice diagrams and situations that are slightly more geometric or visual. Have a look at my n-Lab page: <a href="https://ncatlab.org/timporter/show/simplicial+foundations+for+homotopy+coherence" rel="nofollow noreferrer">https://ncatlab.org/timporter/show/simplicial+foundations+for+homotopy+coherence</a></p> <p>Other material that includes intuitive approaches include my book with Heiner Kamps: Abstract Homotopy and Simple Homotopy Theory, World Scientific, 462pp (ISBN 981-02-1602-5)</p> <p>and the Cubo notes: Abstract Homotopy Theory, the interaction of category theory and homotopy theory, (survey article, updated version of lecture notes from summer school course at Bressanone), Cubo, 5 (2003)115-165, 2003)(and here: <a href="https://ncatlab.org/nlab/files/Abstract-Homotopy.pdf" rel="nofollow noreferrer">https://ncatlab.org/nlab/files/Abstract-Homotopy.pdf</a>).</p> <p>Those do NOT go very far but concentrate on the intuition and basic structure. They may be too elementary for you but are intended to be readable. You can also access a version of the crossed menagerie: <a href="https://ncatlab.org/timporter/show/crossed+menagerie" rel="nofollow noreferrer">https://ncatlab.org/timporter/show/crossed+menagerie</a> which discusses a lot of the relationship with non-abelian cohomology, if that is what is of interest to you. </p> <p>I hope this helps.</p> Answer 5: <p>There are now more resources since the question was raised. Markus Land's <a href="https://link.springer.com/book/10.1007/978-3-030-61524-6" rel="nofollow noreferrer"><em>Introduction to infinity-categories</em></a> grew out from lecture notes, and seemingly much better fitted for pedagogy.</p> <p>On the other hands, due (or thanks?) to pandemic, there are more video resources. For example, there is a series of lectures by Achim Krause and Thomas Nikolaus <a href="https://www.youtube.com/playlist?list=PLsmqTkj4MGTDenpj574aSvIRBROwCugoB" rel="nofollow noreferrer">recorded and uploaded on the YouTube channel</a>.</p> Answer 6: <p>I found <a href="https://arxiv.org/abs/1303.4669" rel="nofollow noreferrer">A Whirlwind Tour of the World of (∞,1)-categories</a> by Omar Antolín Camarena (a student of Jacob Lurie) to be quite insightful for quasicategories.</p> Answer 7: <p>Here are some extra references. (Maybe this is easier to read: <a href="https://pastebin.com/raw/se05pF8E" rel="noreferrer">link</a>)</p> <pre><code>├── Complete Segal Spaces │   ├── [BC, 10 Pages] Equivariant Complete Segal Spaces.pdf │   ├── [Brito, 26 Pages] Segal Objects and the Grothendieck Construction.pdf │   ├── [Osborne, 25 Pages] Decomposing the Classifying Diagram in Terms of Classifying Spaces of Groups.pdf │   ├── [Rasekh, 17 Pages] A Model for the Higher Category of Higher Categories.pdf │   ├── [Rasekh, 44 Pages] Complete Segal Objects.pdf │   ├── [Rasekh, 478 Pages] A Theory of Elementary Higher Toposes.pdf │   ├── [Rasekh, 61 Pages] Introduction to Complete Segal Spaces.pdf │   ├── [Rasekh, 75 Pages] Yoneda Lemma for Simplicial Spaces.pdf │   ├── [Rasekh, 81 Pages] Cartesian Fibrations and Representability.pdf │   ├── [Stenzel, 21 Pages] Univalence and Completeness of Segal Objects.pdf │   └── [Stenzel, 26 Pages] Bousfield-Segal Spaces.pdf ├── Elementary ∞-Topoi │   ├── [Rasekh, 10 Pages] Yoneda Lemma for Elementary Higher Toposes.pdf │   ├── [Rasekh, 30 Pages] Filter Quotients and Non-Presentable (∞,1)-Toposes.pdf │   ├── [Rasekh, 40 Pages] A Theory of Elementary Higher Toposes.pdf │   ├── [Rasekh, 51 Pages] Every Elementary Higher Topos Has a Natural Number Object.pdf │   └── [Rasekh, 84 Pages] An Elementary Approach to Truncations.pdf ├── Enriched (∞,1)-Categories │   ├── [AMR, 68 Pages] Factorization Homology of Enriched ∞-Categories.pdf │   ├── [GH, 100 Pages] Enriched ∞-Categories via Non-Symmetric ∞-Operads.pdf │   ├── [Haugseng, 29 Pages] Bimodules and Natural Transformations for Enriched ∞-Categories.pdf │   └── [Haugseng, 52 Pages] Rectification of Enriched Infinity-Categories.pdf ├── General References │   ├── [AL, 26 Pages] Exponentiable Higher Toposes.pdf │   ├── [Arctaedius, 38 Pages] Grothendieck's Homotopy Hypothesis and the Homotopy Theory of Homotopy Theories.pdf │   ├── [Bergner, 13 Pages] A Survey of (∞, 1)-Categories.pdf │   ├── [Bergner, 287 Pages] The Homotopy Theory of (∞,1)-Categories.pdf │   ├── [Bergner, 29 Pages] A Survey of Models for (∞,n)-Categories.pdf │   ├── [Bergner, 39 Pages] Workshop on the Homotopy Theory of Homotopy Theories.pdf │   ├── [Camarena, 45 Pages] A Whirlwind Tour of the World of (∞,1)-Categories.pdf │   ├── [Clough, 35 Pages] An Outline of the Theory of (∞,1)-Categories.pdf │   ├── [Dorn, 99 Pages] Basic concepts in homotopy theory.pdf │   ├── [HF, 148 Pages] The Homotopy Theory of (∞,1)-Categories.pdf │   ├── [Mazel-Gee, 26 Pages] The Zen of ∞-Categories.pdf │   ├── [Porter, 37 Pages] 𝒮-categories, 𝒮-groupoids, Segal categories and quasicategories.pdf │   ├── [Porter, 54 Pages] Spaces as ∞-groupoids.pdf │   ├── [Porter, 759 Pages] The Crossed Menagerie.pdf │   ├── [Schommer-Pries, 65 Pages] Dualizability in Low-Dimensional Higher Category Theory.pdf │   └── [Simpson, 653 Pages] Homotopy Theory of Higher Categories.pdf ├── Model (∞,1)-Categories │   ├── [LM, 21 Pages] From Fractions to Complete Segal Spaces.pdf │   ├── [Mazel-Gee, 16 Pages] A User's Guide to Co⧸Cartesian Fibrations.pdf │   ├── [Mazel-Gee, 20 Pages] Quillen Adjunctions Induce Adjunctions of Quasicategories.pdf │   ├── [Mazel-Gee, 26 Pages] The Universality of the Rezk Nerve.pdf │   ├── [Mazel-Gee, 29 Pages] Model ∞-Categories II: Quillen Adjunctions.pdf │   ├── [Mazel-Gee, 34 Pages] Model ∞-Categories III: The Fundamental Theorem.pdf │   ├── [Mazel-Gee, 41 Pages] All About the Grothendieck Construction.pdf │   ├── [Mazel-Gee, 43 Pages] Hammocks and Fractions in Relative ∞-Categories.pdf │   ├── [Mazel-Gee, 528 Slides] Goerss–Hopkins obstruction theory for ∞-Categories.pdf │   ├── [Mazel-Gee, 545 Pages] Goerss–Hopkins obstruction theory via model ∞-categories.pdf │   ├── [Mazel-Gee, 54 Pages] Goerss–Hopkins Obstruction Theory for ∞-Categories.pdf │   ├── [Mazel-Gee, 66 Pages] Model ∞-Categories I: Some Pleasant Properties of the ∞-Category of Simplicial Spaces.pdf │   └── [Mazel-Gee, 6 Pages] 𝔼_∞ Automorphisms of Motivic Morava E-Theories.pdf ├── Other Models │   ├── Batanin ∞-Categories │   │   └── [Ara, 168 Pages] Sur les ∞-groupoïdes de Grothendieck et une variante ∞-catégorique.pdf │   ├── Grothendieck–Maltsiniotis ∞-Categories │   │   ├── [AL, 65 Pages] The Folk Model Category Structure on Strict ω-Categories Is Monoidal.pdf │   │   ├── [AM, 123 Pages] A Quillen's Theorem A for Strict ∞-Categories II: The ∞-Categorical Proof.pdf │   │   ├── [AM, 227 Pages] Join and Slices for Strict ∞-Categories.pdf │   │   ├── [AM, 25 Pages] The Brown–Golasinski Model Structure on Strict ∞-Groupoids Revisited.pdf │   │   ├── [AM, 42 Pages] Comparison of the n-Categorical Nerves.pdf │   │   ├── [AM, 51 Pages] A Quillen's Theorem A for Strict ∞-Categories I: The Simplicial Proof.pdf │   │   ├── [AM, 68 Pages] The Homotopy Type of the ∞-Category Associated to a Simplicial Complex.pdf │   │   ├── [AM, 92 Pages] Towards a Thomason Model Structure on the Category of Strict n-Categories.pdf │   │   ├── [Ara, 22 Pages] On Homotopy Types Modelized by Strict ∞-Groupoids.pdf │   │   ├── [Ara, 22 Pages] Strict ∞-Groupoids Are Grothendieck ∞-Groupoids.pdf │   │   ├── [Ara, 33 Pages] A Quillen Theorem B for Strict ∞-Categories.pdf │   │   ├── [Ara, 41 Pages] The Groupoidal Analogue Θ to Joyal's Category Θ Is a Test Category.pdf │   │   └── [Ara, 58 Pages] On the Homotopy Theory of Grothendieck ∞-Groupoids.pdf │   ├── Miscellany │   │   ├── [CL, 61 Pages] Weak ∞-Categories via Terminal Coalgebras.pdf │   │   ├── [Harpaz, 69 Pages] Quasi-Unital ∞-Categories.pdf │   │   ├── [LM, 68 Pages] Linear Quasi-Categories as Templicial Modules.pdf │   │   └── [Nikolaus, 26 Pages] Algebraic Models for Higher Categories.pdf │   ├── Relative Categories │   │   ├── [BK, 19 Pages] n-Relative Categories.pdf │   │   ├── [BK, 27 Pages] Relative Categories: Another Model for the Homotopy Theory of Homotopy Theories.pdf │   │   ├── [BK, 3 Pages] A Thomason-Like Quillen Equivalence Between Quasi-Categories and Relative Categories.pdf │   │   ├── [BK, 5 Pages] In the Category of Relative Categories the Rezk Equivalences Are Exactly the DK-equivalences.pdf │   │   └── [Meier, 21 Pages] Fibration Categories Are Fibrant Relative Categories.pdf │   └── Topological Categories │   ├── [Amrani, 22 Pages] A Model Structure on the Category of Topological Categories.pdf ├── Parametrised (∞,1)-Categories │   ├── [BDGNS, 11 Pages] Parametrized Higher Category Theory and Higher Algebra: A General Introduction.pdf │   ├── [BDGNS, 23 Pages] Parametrized Higher Category Theory and Higher Algebra: Exposé I -- Elements of Parametrized Higher Category Theory.pdf │   ├── [Nardin, 21 Pages] Parametrized Higher Category Theory and Higher Algebra: Exposé IV -- Stability With Respect to an Orbital ∞-Category.pdf │   ├── [Shah, 81 Pages] Parametrized Higher Category Theory and Higher Algebra: Exposé II - Indexed Homotopy Limits and Colimits.pdf │   └── [Shah, 86 Slides] Parametrized Higher Category Theory.pdf ├── Quasicategories │   ├── [AL, 26 Pages] Exponentiable Higher Toposes.pdf │   ├── [Beardsley, 78 Pages] Coalgebraic Structure and Intermediate Hopf–Galois Extensions of Thom Spectra in Quasicategories.pdf │   ├── [Berman, 13 Pages] On Lax Limits in ∞-Categories.pdf │   ├── [BG, 11 Pages] On the Fibrewise Effective Burnside ∞-Category.pdf │   ├── [BM, 30 Pages] Spectral Sequences in (∞,1)-Categories.pdf │   ├── [BS, 18 Pages] Fibrations in ∞-Category Theory.pdf │   ├── [BV, 267 Pages] Homotopy Invariant Algebraic Structures On Topological Spaces.djvu │   ├── [Campbell, 3 Pages] A Counterexample in Quasi-Category Theory.pdf │   ├── [Cisinski, 204 Pages] Algèbre Homotopique et Catégories Supérieures.pdf │   ├── [Cisinski, 446 Pages] Higher Categories and Homotopical Algebra.pdf │   ├── [Cisisnki, 57 Pages] Catégories Supérieures et Théorie des Topos.pdf │   ├── [DS, 46 Pages] Mapping Spaces in Quasi-Categories.pdf │   ├── [Fiore, 24 Pages] Quasicategorical Adjunctions.pdf │   ├── [GR, 11 Pages] Simplified HTT 4.3.2.15.pdf │   ├── [GR, 85 Pages] Some Higher Algebra.pdf │   ├── [Gregoric, 21 Pages] Gregoric Blitzkrieg.pdf │   ├── [Groth, 77 Pages] A Short Course on ∞-Categories.pdf │   ├── [Harpaz, 116 Pages] Little Cube Algebras and Factorisation Homology.pdf │   ├── [Harpaz, 7 Pages] Limits, colimits and adjunctions in ∞-categories.pdf │   ├── [Haugseng, 12 Pages] On (Co)ends in ∞-Categories.pdf │   ├── [Haugseng, 28 Pages] A Fibrational Mate Correspondence for ∞-Categories.pdf │   ├── [Haugseng, 87 Pages] Introduction to ∞-Categories.pdf │   ├── [HH, 684 Pages] Higher Categories I &amp; II │   │   ├── 0 [Hebestreit, 13 Pages] A Fairytale.pdf │   │   ├── 10 [Hebestreit, 11 Pages] Fat and Thin Slices.pdf │   │   ├── 11 [Hebestreit, 43 Pages] Cartesian Fibrations.pdf │   │   ├── 12 [Hebestreit, 121 Pages] Straightening and Unstraightening.pdf │   │   ├── 12' [HH, 53 Pages] Straightening and Unstraightening (Heuts's Notes).pdf │   │   ├── 13 [Hebestreit, 14 Pages] Homotopy Colimits.pdf │   │   ├── 14 [Hebestreit, 10 Pages] Simplicial Model Categories.pdf │   │   ├── 15 [Hebestreit, 35 Pages] Yoneda's Lemma, Adjunctions and (Co)Limits.pdf │   │   ├── 1 [Hebestreit, 37 Pages] Categories.pdf │   │   ├── 2 [Hebestreit, 25 Pages] Simplicial Sets.pdf │   │   ├── 3 [Hebestreit, 10 Pages] Quasicategories.pdf │   │   ├── 4 [Hebestreit, 53 Pages] Simplicial Categories.pdf │   │   ├── 5 [Hebestreit, 38 Pages] Simplicial Homotopy Theory.pdf │   │   ├── 6 [Hebestreit, 57 Pages] Quasicategories and Anima.pdf │   │   ├── 7 [Hebestreit, 67 Pages] Equivalences, Equivalences, Equivalences.pdf │   │   ├── 8 [Hebestreit, 9 Pages] A Fairytale.pdf │   │   └── 9 [Hebestreit, 88 Pages] Localisations and Model Categories.pdf │   ├── [Hinich, 111 Pages] Lectures on ∞-Categories.pdf │   ├── [Joyal, 244 Pages] Notes on Quasi-Categories.pdf │   ├── [Joyal, 350 Pages] The Theory of Quasi-Categories and its Applications II.pdf │   ├── [Joyal, 479 Pages] The Theory of Quasi-Categories.pdf │   ├── [Lejay, 139 Pages] Algèbres à Factorisation et Topos Supérieurs Exponentiables.pdf │   ├── [Lurie, 60 Pages] On ∞-Topoi.pdf │   ├── [Lurie, 841 Pages] Kerodon.pdf │   ├── [Lurie, 949 Pages] Higher Topos Theory.pdf │   ├── [Lysenko, 208 Pages] Lysenko's comments to Gaitsgory–Rozenblyum.pdf │   ├── [Morel, 118 Pages] ∞-Categories.pdf │   ├── [Nguyen, 94 Pages] Theorems in Higher Category Theory and Applications.pdf │   ├── [NRS, 21 Pages] Adjoint Functor Theorems for ∞-Categories.pdf │   ├── [Porta, 100 Pages] Derived formal moduli problems.pdf │   ├── [Rezk, 11 Pages] Degenerate Edges of Cartesian Fibrations are Cartesian Edges.pdf │   ├── [Rezk, 175 Pages] Stuff About Quasicategories.pdf │   ├── [Rezk, 50 Pages] Toposes and homotopy toposes.pdf │   ├── [Riehl, 20 Pages] Quasi-Categories as (∞,1)-Categories.pdf │   ├── [Riehl, 292 Pages] Categorical homotopy theory.pdf │   ├── [Riehl, 9 Pages] Associativity Data in an (∞,1)-Category.pdf │   ├── [Rovelli, 38 Pages] Weighted Limits in an (∞,1)-Category.pdf │   ├── [RS, 151 Pages] Notes on Higher Categories.pdf │   ├── [RV, 33 Pages] Completeness Results for Quasi-Categories of Algebras, Homotopy Limits, and Related General Constructions.pdf │   ├── [Stevenson, 12 Pages] Stability for Inner Fibrations Revisited.pdf │   ├── [Tanaka, 14 Pages] Functors (Between ∞-Categories) That Aren't Strictly Unital.pdf │   ├── [Thanh, 46 Pages] Quasicategories.pdf │   ├── [Wong, 83 Pages] The Grothendieck construction in enriched, internal and ∞-Category Theory.pdf │   ├── [Zsámboki, 31 Pages] A summary of higher topos theory.pdf │   └── [無, 5 Pages] HTT ToC.pdf ├── Simplicial Categories │   ├── [Bergner, 16 Pages] A model category structure on the category of simplicial categories.pdf │   ├── [Bergner, 22 Pages] Complete Segal Spaces Arising From Simplicial Categories.pdf │   ├── [Bergner, 40 Pages] Three models for the homotopy theory of homotopy theories.pdf │   ├── [Cordier, 21 Pages] Sur la notion de diagramme homotopiquement cohérent.pdf │   ├── [CP, 26 Pages] Vogt's Theorem on Categories of Homotopy Coherent Diagrams.pdf │   ├── [CP, 54 Pages] Homotopy Coherent Category Theory.pdf │   ├── [DS, 29 Pages] Rigidification of Quasi-Categories.pdf │   ├── [DS, 64 Pages] Mapping Spaces in Quasi-Categories.pdf │   ├── [Hinich, 23 Pages] Homotopy Coherent Nerve in Deformation Theory.pdf │   ├── [HK, 14 Pages] Mapping Spaces in Homotopy Coherent Nerves.pdf │   ├── [Joyal, 66 Pages] Quasi-Categories vs Simplicial Categories.pdf │   ├── [Riehl, 16 Pages] On the Structure of Simplicial Categories Associated to Quasi-Categories.pdf │   ├── [Riehl, 26 Pages] Homotopy Coherent Structures.pdf │   ├── [Riehl, 292 Pages] Categorical Homotopy Theory.pdf │   └── [Riehl, 7 Pages] Understanding the Homotopy Coherent Nerve.pdf ├── Unicity │   ├── [Bergner, 16 Pages] Equivalence of Models for Equivariant (∞,1)-Categories.pdf │   ├── [BS, 47 Pages] On the Unicity of the Homotopy Theory of Higher Categories.pdf │   ├── [JT, 49 Pages] Quasi-Categories vs Segal Spaces.pdf │   ├── [Riehl, 58 Pages] Seminar Notes on the Barwick–Schommer–Pries Unicity Theorem.pdf │   └── [Toën, 32 Pages] Vers une axiomatisation de la théorie des catégories supérieures.pdf └── Un⧸Straightening ├── [AF, 89 Pages] Fibrations of ∞-Categories.pdf ├── [BGN, 19 Pages] Dualizing Cartesian and Cocartesian Fibrations.pdf ├── [BS, 18 Pages] Fibrations in ∞-Category Theory.pdf ├── [Campbell, 31 Slides] A modular proof of the straightening theorem.pdf ├── [GHN, 42 Pages] Lax Colimits and Free Fibrations in ∞-Categories.pdf ├── [GR, 85 Pages] Some Higher Algebra.pdf ├── [Hebestreit, 121 Pages] Straightening and Unstraightening.pdf ├── [HH, 53 Pages] Straightening and Unstraightening (Heuts's Notes).pdf ├── [HM, 17 Pages] Left Fibrations and Homotopy Colimits II.pdf ├── [HM, 27 Pages] Left Fibrations and Homotopy Colimits.pdf ├── [Mazel-Gee, 16 Pages] A User's Guide to (Co)Cartesian Fibrations.pdf ├── [Mazel-Gee, 41 Pages] All About the Grothendieck Construction.pdf ├── [Noel, 1 Page] Cartesian Model Structure.pdf ├── [PK, 43 Pages] Straightening and Unstraightening.pdf ├── [Richardson, 14 Pages] Mapping Spaces and Straightening-Unstraightening.pdf ├── [Ruit, 71 Pages] Grothendieck constructions in higher category theory.pdf ├── [Stevenson, 41 Pages] Model Structures for Correspondences and Bifibrations.pdf ├── [Stevenson, 49 Pages] Covariant Model Structures and Simplicial Localization.pdf └── [Wong, 83 Pages] The Grothendieck Constructionin Enriched, Internal and ∞-Category Theory.pdf </code></pre> Answer 8: <p>Moritz Groth put up some excellent lecture notes: <a href="https://arxiv.org/abs/1007.2925v2" rel="nofollow noreferrer">https://arxiv.org/abs/1007.2925v2</a></p> <p>If you are more categorically minded, Emily Riehl's book has a lot about quasi-categories: <a href="http://www.math.jhu.edu/%7Eeriehl/cathtpy.pdf" rel="nofollow noreferrer">http://www.math.jhu.edu/~eriehl/cathtpy.pdf</a> and her website has a lot of exposition about cosmoi, an approach to <span class="math-container">$\infty$</span>-categories via 2-category theory, and not requiring one to work in the setting of quasi-categories: <a href="http://www.math.jhu.edu/%7Eeriehl/" rel="nofollow noreferrer">http://www.math.jhu.edu/~eriehl/</a></p> Answer 9: <p>It depends on from what direction you are coming and in what direction you are wanting to go! </p> <p>As one of the developers of the basic theory from way back, I like to approach things via homotopy coherence as that is where the quasi-category approach comes from, and it also more or less guarantees some nice diagrams and situations that are slightly more geometric or visual. Have a look at my n-Lab page: <a href="https://ncatlab.org/timporter/show/simplicial+foundations+for+homotopy+coherence" rel="nofollow noreferrer">https://ncatlab.org/timporter/show/simplicial+foundations+for+homotopy+coherence</a></p> <p>Other material that includes intuitive approaches include my book with Heiner Kamps: Abstract Homotopy and Simple Homotopy Theory, World Scientific, 462pp (ISBN 981-02-1602-5)</p> <p>and the Cubo notes: Abstract Homotopy Theory, the interaction of category theory and homotopy theory, (survey article, updated version of lecture notes from summer school course at Bressanone), Cubo, 5 (2003)115-165, 2003)(and here: <a href="https://ncatlab.org/nlab/files/Abstract-Homotopy.pdf" rel="nofollow noreferrer">https://ncatlab.org/nlab/files/Abstract-Homotopy.pdf</a>).</p> <p>Those do NOT go very far but concentrate on the intuition and basic structure. They may be too elementary for you but are intended to be readable. You can also access a version of the crossed menagerie: <a href="https://ncatlab.org/timporter/show/crossed+menagerie" rel="nofollow noreferrer">https://ncatlab.org/timporter/show/crossed+menagerie</a> which discusses a lot of the relationship with non-abelian cohomology, if that is what is of interest to you. </p> <p>I hope this helps.</p> Answer 10: <p>There are now more resources since the question was raised. Markus Land's <a href="https://link.springer.com/book/10.1007/978-3-030-61524-6" rel="nofollow noreferrer"><em>Introduction to infinity-categories</em></a> grew out from lecture notes, and seemingly much better fitted for pedagogy.</p> <p>On the other hands, due (or thanks?) to pandemic, there are more video resources. For example, there is a series of lectures by Achim Krause and Thomas Nikolaus <a href="https://www.youtube.com/playlist?list=PLsmqTkj4MGTDenpj574aSvIRBROwCugoB" rel="nofollow noreferrer">recorded and uploaded on the YouTube channel</a>.</p>
Errata for Linear algebraic groups by Springer
259,138
6
https://mathoverflow.net/questions/259138/errata-for-linear-algebraic-groups-by-springer
<p>Does any one prepared a list of errata for Linear algebraic groups by Springer.</p> <p>I could not find any in Google search.</p> <p>First typo that i came across is in page 6, Regular functions and ringed spaces:</p> <blockquote> <p>If $U $ and $V$ are open subsets and $U\subset V$, restriction defines a $k$ - algebra homomorphism $\mathcal{O}(U) \rightarrow \mathcal{O}(V)$.</p> </blockquote> <p>I believe it has to be </p> <blockquote> <p>If $U $ and $V$ are open subsets and $U\subset V$, restriction defines a $k$ - algebra homomorphism $\mathcal{O}(V) \rightarrow \mathcal{O}(U)$.</p> </blockquote> <p>Am i misunderstanding something or is it really a typo.</p> <p>I would be grateful if some one can provide list of typos in that book.</p>
[ { "answer_id": 259198, "score": 13, "text": "<p>The answer to your specific question is that it's really stated backwards. More generally, your question about lists of errata comes up fairly often here and is hard to answer in detail. It's a legitimate question to ask when looking at relatively advanced books in mathematics. (Maybe a special tag is needed?) But unfortunately, queries to both authors and publishers are apt to be ignored. </p>\n\n<p>As noted in the comments, Tonny Springer himself is no longer alive. His years were 1926-2011, and his death followed a sudden aneurysm. He was a gifted and creative Dutch mathematician whose work has continued to be highly influential (as seen for instance in the title of a well-attended workshop here at UMass in October 2015). His 1978 Notre Dame lectures on linear algebraic groups led to his first edition (Birkhauser, 1981). But in attempting to avoid some of the tricky aspects of algebraic geometry in prime characteristic, he didn't quite succeed at first. He also found a need to add some topics to the book, so a second edition followed in 1998. I'm assuming this is the edition you are looking at. </p>\n\n<p>[By the way, the publisher Birkhauser (with an Umlaut over the a) is the same, but their Boston branch was originally an offshoot of the Swiss publishing house and was later acquired by Springer-Verlag: no relation to T.A. Springer. Indeed, when there was concern decades ago about the confusion of Springer-Verlag with the right-wing Axel Springer publishing company, Tonny Springer posted on his office door in Utrecht a headline <em>Springer $\\neq$ Springer</em> from the Springer-Verlag disclaimer published in mathematical journals at the time.]</p>\n\n<p>Concerning errata in books, these very often occur but are inadequately tracked by publishers or living authors. An exception is the American Mathematical Society, which offers to authors an online bookpage where errata and other supplementary materials can be posted. I've maintained similar lists on my own webpage, never having tried out my own books on actual students.</p>\n\n<p>Errata come in all sizes, ranging from obvious misprints to mistaken assertions to exercises poorly stated or misplaced. Those I've noticed and marked in my own copy of Springer's second edition tend to be minor, but I try to keep a list of page numbers. It's amusing to spot the errors (sometimes more than one) on his pages 195, 218, 265, 284, 304, 325, 332. Probably the most serious oversight is on page 321, where one index is apparently inexact. There are also some items such as $F$-<em>reductive</em> missing from the list of terminology at the end of the book. </p>\n" }, { "answer_id": 311762, "score": 1, "text": "<p>Here is a different one, on page 299: Proposition 17.3.13(iii) giving the relative root system for quasi-simple groups of type <span class=\"math-container\">$D_{n}$</span> is not correct as stated. I believe the following is correct, but please comment if not!</p>\n\n<p>If <span class=\"math-container\">$n &gt; rd$</span>, then the relative root system has type <span class=\"math-container\">$B_r$</span> if <span class=\"math-container\">$d = 1$</span> and type <span class=\"math-container\">$BC_r$</span> if <span class=\"math-container\">$d &gt; 1$</span>. If <span class=\"math-container\">$n = rd$</span>, then the relative root system has type <span class=\"math-container\">$D_r$</span> if <span class=\"math-container\">$d = 1$</span> and type <span class=\"math-container\">$C_r$</span> if <span class=\"math-container\">$d &gt; 1$</span>. The case <span class=\"math-container\">$n = rd$</span> only occurs when the group is an inner form (type <span class=\"math-container\">${}^1 D_{n,r}^{(d)}$</span> in Tits' notation).</p>\n\n<p>(I don't know a reference. I didn't read this section of Springer, but I worked out the relative root datum from the absolute one in each case. Over a non-archimedean local field the answer agrees with Tits' Corvallis table.)</p>\n\n<p>Updated: also part (i) is incorrect if <span class=\"math-container\">$rd = n-1$</span>. In this case the group is an outer form and the last element <span class=\"math-container\">$a_{rd}$</span> in Springer's list should be replaced by the conjugate pair <span class=\"math-container\">$a_{rd}$</span>, <span class=\"math-container\">$a_{rd+1}$</span> (i.e. <span class=\"math-container\">$a_{n-1}$</span>, <span class=\"math-container\">$a_n$</span>). See Tits' Boulder article from 1965.</p>\n" }, { "answer_id": 259198, "score": 13, "text": "<p>The answer to your specific question is that it's really stated backwards. More generally, your question about lists of errata comes up fairly often here and is hard to answer in detail. It's a legitimate question to ask when looking at relatively advanced books in mathematics. (Maybe a special tag is needed?) But unfortunately, queries to both authors and publishers are apt to be ignored. </p>\n\n<p>As noted in the comments, Tonny Springer himself is no longer alive. His years were 1926-2011, and his death followed a sudden aneurysm. He was a gifted and creative Dutch mathematician whose work has continued to be highly influential (as seen for instance in the title of a well-attended workshop here at UMass in October 2015). His 1978 Notre Dame lectures on linear algebraic groups led to his first edition (Birkhauser, 1981). But in attempting to avoid some of the tricky aspects of algebraic geometry in prime characteristic, he didn't quite succeed at first. He also found a need to add some topics to the book, so a second edition followed in 1998. I'm assuming this is the edition you are looking at. </p>\n\n<p>[By the way, the publisher Birkhauser (with an Umlaut over the a) is the same, but their Boston branch was originally an offshoot of the Swiss publishing house and was later acquired by Springer-Verlag: no relation to T.A. Springer. Indeed, when there was concern decades ago about the confusion of Springer-Verlag with the right-wing Axel Springer publishing company, Tonny Springer posted on his office door in Utrecht a headline <em>Springer $\\neq$ Springer</em> from the Springer-Verlag disclaimer published in mathematical journals at the time.]</p>\n\n<p>Concerning errata in books, these very often occur but are inadequately tracked by publishers or living authors. An exception is the American Mathematical Society, which offers to authors an online bookpage where errata and other supplementary materials can be posted. I've maintained similar lists on my own webpage, never having tried out my own books on actual students.</p>\n\n<p>Errata come in all sizes, ranging from obvious misprints to mistaken assertions to exercises poorly stated or misplaced. Those I've noticed and marked in my own copy of Springer's second edition tend to be minor, but I try to keep a list of page numbers. It's amusing to spot the errors (sometimes more than one) on his pages 195, 218, 265, 284, 304, 325, 332. Probably the most serious oversight is on page 321, where one index is apparently inexact. There are also some items such as $F$-<em>reductive</em> missing from the list of terminology at the end of the book. </p>\n" }, { "answer_id": 311762, "score": 1, "text": "<p>Here is a different one, on page 299: Proposition 17.3.13(iii) giving the relative root system for quasi-simple groups of type <span class=\"math-container\">$D_{n}$</span> is not correct as stated. I believe the following is correct, but please comment if not!</p>\n\n<p>If <span class=\"math-container\">$n &gt; rd$</span>, then the relative root system has type <span class=\"math-container\">$B_r$</span> if <span class=\"math-container\">$d = 1$</span> and type <span class=\"math-container\">$BC_r$</span> if <span class=\"math-container\">$d &gt; 1$</span>. If <span class=\"math-container\">$n = rd$</span>, then the relative root system has type <span class=\"math-container\">$D_r$</span> if <span class=\"math-container\">$d = 1$</span> and type <span class=\"math-container\">$C_r$</span> if <span class=\"math-container\">$d &gt; 1$</span>. The case <span class=\"math-container\">$n = rd$</span> only occurs when the group is an inner form (type <span class=\"math-container\">${}^1 D_{n,r}^{(d)}$</span> in Tits' notation).</p>\n\n<p>(I don't know a reference. I didn't read this section of Springer, but I worked out the relative root datum from the absolute one in each case. Over a non-archimedean local field the answer agrees with Tits' Corvallis table.)</p>\n\n<p>Updated: also part (i) is incorrect if <span class=\"math-container\">$rd = n-1$</span>. In this case the group is an outer form and the last element <span class=\"math-container\">$a_{rd}$</span> in Springer's list should be replaced by the conjugate pair <span class=\"math-container\">$a_{rd}$</span>, <span class=\"math-container\">$a_{rd+1}$</span> (i.e. <span class=\"math-container\">$a_{n-1}$</span>, <span class=\"math-container\">$a_n$</span>). See Tits' Boulder article from 1965.</p>\n" } ]
Question: Errata for Linear algebraic groups by Springer Answer 1: <p>The answer to your specific question is that it's really stated backwards. More generally, your question about lists of errata comes up fairly often here and is hard to answer in detail. It's a legitimate question to ask when looking at relatively advanced books in mathematics. (Maybe a special tag is needed?) But unfortunately, queries to both authors and publishers are apt to be ignored. </p> <p>As noted in the comments, Tonny Springer himself is no longer alive. His years were 1926-2011, and his death followed a sudden aneurysm. He was a gifted and creative Dutch mathematician whose work has continued to be highly influential (as seen for instance in the title of a well-attended workshop here at UMass in October 2015). His 1978 Notre Dame lectures on linear algebraic groups led to his first edition (Birkhauser, 1981). But in attempting to avoid some of the tricky aspects of algebraic geometry in prime characteristic, he didn't quite succeed at first. He also found a need to add some topics to the book, so a second edition followed in 1998. I'm assuming this is the edition you are looking at. </p> <p>[By the way, the publisher Birkhauser (with an Umlaut over the a) is the same, but their Boston branch was originally an offshoot of the Swiss publishing house and was later acquired by Springer-Verlag: no relation to T.A. Springer. Indeed, when there was concern decades ago about the confusion of Springer-Verlag with the right-wing Axel Springer publishing company, Tonny Springer posted on his office door in Utrecht a headline <em>Springer $\neq$ Springer</em> from the Springer-Verlag disclaimer published in mathematical journals at the time.]</p> <p>Concerning errata in books, these very often occur but are inadequately tracked by publishers or living authors. An exception is the American Mathematical Society, which offers to authors an online bookpage where errata and other supplementary materials can be posted. I've maintained similar lists on my own webpage, never having tried out my own books on actual students.</p> <p>Errata come in all sizes, ranging from obvious misprints to mistaken assertions to exercises poorly stated or misplaced. Those I've noticed and marked in my own copy of Springer's second edition tend to be minor, but I try to keep a list of page numbers. It's amusing to spot the errors (sometimes more than one) on his pages 195, 218, 265, 284, 304, 325, 332. Probably the most serious oversight is on page 321, where one index is apparently inexact. There are also some items such as $F$-<em>reductive</em> missing from the list of terminology at the end of the book. </p> Answer 2: <p>Here is a different one, on page 299: Proposition 17.3.13(iii) giving the relative root system for quasi-simple groups of type <span class="math-container">$D_{n}$</span> is not correct as stated. I believe the following is correct, but please comment if not!</p> <p>If <span class="math-container">$n &gt; rd$</span>, then the relative root system has type <span class="math-container">$B_r$</span> if <span class="math-container">$d = 1$</span> and type <span class="math-container">$BC_r$</span> if <span class="math-container">$d &gt; 1$</span>. If <span class="math-container">$n = rd$</span>, then the relative root system has type <span class="math-container">$D_r$</span> if <span class="math-container">$d = 1$</span> and type <span class="math-container">$C_r$</span> if <span class="math-container">$d &gt; 1$</span>. The case <span class="math-container">$n = rd$</span> only occurs when the group is an inner form (type <span class="math-container">${}^1 D_{n,r}^{(d)}$</span> in Tits' notation).</p> <p>(I don't know a reference. I didn't read this section of Springer, but I worked out the relative root datum from the absolute one in each case. Over a non-archimedean local field the answer agrees with Tits' Corvallis table.)</p> <p>Updated: also part (i) is incorrect if <span class="math-container">$rd = n-1$</span>. In this case the group is an outer form and the last element <span class="math-container">$a_{rd}$</span> in Springer's list should be replaced by the conjugate pair <span class="math-container">$a_{rd}$</span>, <span class="math-container">$a_{rd+1}$</span> (i.e. <span class="math-container">$a_{n-1}$</span>, <span class="math-container">$a_n$</span>). See Tits' Boulder article from 1965.</p> Answer 3: <p>The answer to your specific question is that it's really stated backwards. More generally, your question about lists of errata comes up fairly often here and is hard to answer in detail. It's a legitimate question to ask when looking at relatively advanced books in mathematics. (Maybe a special tag is needed?) But unfortunately, queries to both authors and publishers are apt to be ignored. </p> <p>As noted in the comments, Tonny Springer himself is no longer alive. His years were 1926-2011, and his death followed a sudden aneurysm. He was a gifted and creative Dutch mathematician whose work has continued to be highly influential (as seen for instance in the title of a well-attended workshop here at UMass in October 2015). His 1978 Notre Dame lectures on linear algebraic groups led to his first edition (Birkhauser, 1981). But in attempting to avoid some of the tricky aspects of algebraic geometry in prime characteristic, he didn't quite succeed at first. He also found a need to add some topics to the book, so a second edition followed in 1998. I'm assuming this is the edition you are looking at. </p> <p>[By the way, the publisher Birkhauser (with an Umlaut over the a) is the same, but their Boston branch was originally an offshoot of the Swiss publishing house and was later acquired by Springer-Verlag: no relation to T.A. Springer. Indeed, when there was concern decades ago about the confusion of Springer-Verlag with the right-wing Axel Springer publishing company, Tonny Springer posted on his office door in Utrecht a headline <em>Springer $\neq$ Springer</em> from the Springer-Verlag disclaimer published in mathematical journals at the time.]</p> <p>Concerning errata in books, these very often occur but are inadequately tracked by publishers or living authors. An exception is the American Mathematical Society, which offers to authors an online bookpage where errata and other supplementary materials can be posted. I've maintained similar lists on my own webpage, never having tried out my own books on actual students.</p> <p>Errata come in all sizes, ranging from obvious misprints to mistaken assertions to exercises poorly stated or misplaced. Those I've noticed and marked in my own copy of Springer's second edition tend to be minor, but I try to keep a list of page numbers. It's amusing to spot the errors (sometimes more than one) on his pages 195, 218, 265, 284, 304, 325, 332. Probably the most serious oversight is on page 321, where one index is apparently inexact. There are also some items such as $F$-<em>reductive</em> missing from the list of terminology at the end of the book. </p> Answer 4: <p>Here is a different one, on page 299: Proposition 17.3.13(iii) giving the relative root system for quasi-simple groups of type <span class="math-container">$D_{n}$</span> is not correct as stated. I believe the following is correct, but please comment if not!</p> <p>If <span class="math-container">$n &gt; rd$</span>, then the relative root system has type <span class="math-container">$B_r$</span> if <span class="math-container">$d = 1$</span> and type <span class="math-container">$BC_r$</span> if <span class="math-container">$d &gt; 1$</span>. If <span class="math-container">$n = rd$</span>, then the relative root system has type <span class="math-container">$D_r$</span> if <span class="math-container">$d = 1$</span> and type <span class="math-container">$C_r$</span> if <span class="math-container">$d &gt; 1$</span>. The case <span class="math-container">$n = rd$</span> only occurs when the group is an inner form (type <span class="math-container">${}^1 D_{n,r}^{(d)}$</span> in Tits' notation).</p> <p>(I don't know a reference. I didn't read this section of Springer, but I worked out the relative root datum from the absolute one in each case. Over a non-archimedean local field the answer agrees with Tits' Corvallis table.)</p> <p>Updated: also part (i) is incorrect if <span class="math-container">$rd = n-1$</span>. In this case the group is an outer form and the last element <span class="math-container">$a_{rd}$</span> in Springer's list should be replaced by the conjugate pair <span class="math-container">$a_{rd}$</span>, <span class="math-container">$a_{rd+1}$</span> (i.e. <span class="math-container">$a_{n-1}$</span>, <span class="math-container">$a_n$</span>). See Tits' Boulder article from 1965.</p>
12th grade - Ramanujan Partition theory
259,297
6
https://mathoverflow.net/questions/259297/12th-grade-ramanujan-partition-theory
<p>I've been really trying to prove Ramanujan Partition theory, and different sources give me different explanations. Can someone please explain how Ramanujan (and Euler) found out the following equation for the number of partitions for a given integer? Any help is appreciated thank you so much!</p> <p>$$P(n)\sim\frac1{4n\sqrt3}\exp(\pi\sqrt{2n/3})$$ <img src="https://i.stack.imgur.com/UEQnv.jpg" alt="enter image description here"></p>
[ { "answer_id": 259298, "score": 3, "text": "<p>There is a proof in <a href=\"https://www.amazon.co.uk/Ramanujan-Lectures-Subjects-Suggested-Publishing/dp/0821820230\" rel=\"nofollow noreferrer\">Ramanujan: Twelve Lectures</a> by Hardy.</p>\n" }, { "answer_id": 259320, "score": 3, "text": "<p>Another reference is Atle Selberg's paper \"<a href=\"http://dx.doi.org/10.1007/BF02838915\" rel=\"nofollow noreferrer\">Reflections around the Ramanujan centenary</a>\". More precisely, the appendix, if I remember. It appears in the book \"<a href=\"https://books.google.com/books?id=TT1T8A94xNcC&amp;pg=PA203\" rel=\"nofollow noreferrer\">Ramanujan: essays and surveys</a>\". My recommendation would be to read that before Hardy, for reasons that Selberg himself explains.</p>\n" }, { "answer_id": 259298, "score": 3, "text": "<p>There is a proof in <a href=\"https://www.amazon.co.uk/Ramanujan-Lectures-Subjects-Suggested-Publishing/dp/0821820230\" rel=\"nofollow noreferrer\">Ramanujan: Twelve Lectures</a> by Hardy.</p>\n" }, { "answer_id": 259320, "score": 3, "text": "<p>Another reference is Atle Selberg's paper \"<a href=\"http://dx.doi.org/10.1007/BF02838915\" rel=\"nofollow noreferrer\">Reflections around the Ramanujan centenary</a>\". More precisely, the appendix, if I remember. It appears in the book \"<a href=\"https://books.google.com/books?id=TT1T8A94xNcC&amp;pg=PA203\" rel=\"nofollow noreferrer\">Ramanujan: essays and surveys</a>\". My recommendation would be to read that before Hardy, for reasons that Selberg himself explains.</p>\n" } ]
Question: 12th grade - Ramanujan Partition theory Answer 1: <p>There is a proof in <a href="https://www.amazon.co.uk/Ramanujan-Lectures-Subjects-Suggested-Publishing/dp/0821820230" rel="nofollow noreferrer">Ramanujan: Twelve Lectures</a> by Hardy.</p> Answer 2: <p>Another reference is Atle Selberg's paper "<a href="http://dx.doi.org/10.1007/BF02838915" rel="nofollow noreferrer">Reflections around the Ramanujan centenary</a>". More precisely, the appendix, if I remember. It appears in the book "<a href="https://books.google.com/books?id=TT1T8A94xNcC&amp;pg=PA203" rel="nofollow noreferrer">Ramanujan: essays and surveys</a>". My recommendation would be to read that before Hardy, for reasons that Selberg himself explains.</p> Answer 3: <p>There is a proof in <a href="https://www.amazon.co.uk/Ramanujan-Lectures-Subjects-Suggested-Publishing/dp/0821820230" rel="nofollow noreferrer">Ramanujan: Twelve Lectures</a> by Hardy.</p> Answer 4: <p>Another reference is Atle Selberg's paper "<a href="http://dx.doi.org/10.1007/BF02838915" rel="nofollow noreferrer">Reflections around the Ramanujan centenary</a>". More precisely, the appendix, if I remember. It appears in the book "<a href="https://books.google.com/books?id=TT1T8A94xNcC&amp;pg=PA203" rel="nofollow noreferrer">Ramanujan: essays and surveys</a>". My recommendation would be to read that before Hardy, for reasons that Selberg himself explains.</p>
Mid-Square with all bits set
259,663
6
https://mathoverflow.net/questions/259663/mid-square-with-all-bits-set
<p>Is there a positive 128-bit integer whose square has all middle bits equal to 1?</p> <p>(The "middle bits" are naturally the 65th bit through the 192nd bit, defining the 1st bit as the least significant bit of the full integer.)</p>
[ { "answer_id": 418082, "score": 7, "text": "<blockquote>\n<p>Is there a positive 128-bit integer whose square has all middle bits equal to 1?</p>\n</blockquote>\n<p><em><strong>YES</strong></em>. One is AAAAAAAAAAAAAAAB5555555555555555<sub>16</sub>, which square is 71C71C71C71C71C7FFFFFFFFFFFFFFFFFFFFFFFFFFFFFFFF8E38E38E38E38E39<sub>16</sub>.</p>\n<hr />\n<p>It's asked if there are integers <span class=\"math-container\">$n$</span> such that <span class=\"math-container\">$0&lt;n&lt;2^{2k}$</span> and <span class=\"math-container\">$(n^2\\bmod2^{3k})\\ge2^{3k}-2^k$</span>, in the case <span class=\"math-container\">$k=128/2$</span>.</p>\n<p>As noted in <a href=\"https://mathoverflow.net/posts/comments/1073379?noredirect=1\">comment by Max Alekseyev</a>, that holds for <span class=\"math-container\">$k\\equiv 4\\pmod6$</span> (thus including the question's <span class=\"math-container\">$k=128/2$</span>), because the integer\n<span class=\"math-container\">$$n=\\frac{2^{2k+1}+2^{k+1}-1}3$$</span>\ndoes the trick. It's representation in base <span class=\"math-container\">$4$</span> consists of <span class=\"math-container\">$k/2-1$</span> digits <span class=\"math-container\">$2$</span>, one digit <span class=\"math-container\">$3$</span>, and <span class=\"math-container\">$k/2$</span> digits <span class=\"math-container\">$1$</span>. It holds\n<span class=\"math-container\">$$n^2=\\frac{8^k-1}{9}2^{k+2}+\\frac{8^{k+1}+1}9$$</span>\nwhich representation in base <span class=\"math-container\">$4$</span> consists of the digit <span class=\"math-container\">$1$</span>; the digits <span class=\"math-container\">$301$</span> repeated <span class=\"math-container\">$(k-4)/6$</span> times; the digit <span class=\"math-container\">$3$</span> repeated <span class=\"math-container\">$1+k$</span> times; the digit <span class=\"math-container\">$2$</span>; the digits <span class=\"math-container\">$032$</span> repeated <span class=\"math-container\">$(k-4)/6$</span> times; the digit <span class=\"math-container\">$1$</span>. It follows the center <span class=\"math-container\">$2k$</span> bits are set, with three more on the left and one more on the right.</p>\n<p>We can't use this to conclude exactly how many <span class=\"math-container\">$n$</span> there are, or/and for other values of <span class=\"math-container\">$k$</span>, or/and for an arbitrary value in the middle of the square.</p>\n<p>Here is a brute force search for <span class=\"math-container\">$k\\le21$</span>:</p>\n<div class=\"s-table-container\">\n<table class=\"s-table\">\n<thead>\n<tr>\n<th style=\"text-align: center;\"><span class=\"math-container\">$k$</span> with a solution</th>\n<th style=\"text-align: center;\">#solution(s)</th>\n<th style=\"text-align: right;\">possible <span class=\"math-container\">$n$</span></th>\n</tr>\n</thead>\n<tbody>\n<tr>\n<td style=\"text-align: center;\">4</td>\n<td style=\"text-align: center;\">1</td>\n<td style=\"text-align: right;\">10110101<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">5</td>\n<td style=\"text-align: center;\">2</td>\n<td style=\"text-align: right;\">0010110101<sub>2</sub><br>0101101010<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">10</td>\n<td style=\"text-align: center;\">2</td>\n<td style=\"text-align: right;\">10101010110101010101<sub>2</sub><br>11100010010001100011<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">11</td>\n<td style=\"text-align: center;\">4</td>\n<td style=\"text-align: right;\">0010101010110101010101<sub>2</sub><br>0011100010010001100011<sub>2</sub><br>0101010101101010101010<sub>2</sub><br>0111000100100011000110<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">12</td>\n<td style=\"text-align: center;\">3</td>\n<td style=\"text-align: right;\">000111000100100011000110<sub>2</sub><br>010010001110001000101011<sub>2</sub><br>110001101110000001111001<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">13</td>\n<td style=\"text-align: center;\">3</td>\n<td style=\"text-align: right;\">00001110001001000110001100<sub>2</sub><br>00110001101110000001111001<sub>2</sub><br>01100011011100000011110010<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">14</td>\n<td style=\"text-align: center;\">1</td>\n<td style=\"text-align: right;\">1010001000110010100010011000<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">16</td>\n<td style=\"text-align: center;\">1</td>\n<td style=\"text-align: right;\">10101010101010110101010101010101<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">17</td>\n<td style=\"text-align: center;\">2</td>\n<td style=\"text-align: right;\">0010101010101010110101010101010101<sub>2</sub><br>0101010101010101101010101010101010<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">19</td>\n<td style=\"text-align: center;\">1</td>\n<td style=\"text-align: right;\">11110010011001000101100100110110000011<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">21</td>\n<td style=\"text-align: center;\">4</td>\n<td style=\"text-align: right;\">010100000101011011101010011000101011000111<sub>2</sub><br>011101100111011110011000000000010111000101<sub>2</sub><br>101000001010110111010100110001010110001110<sub>2</sub><br>111101110010111001100001001101110110100001<sub>2</sub></td>\n</tr>\n</tbody>\n</table>\n</div>\n<p>We see and could prove that if <span class=\"math-container\">$n$</span> is a solution for <span class=\"math-container\">$k=k_a+1$</span> with <span class=\"math-container\">$n&lt;2^{2k_a}$</span>, then <span class=\"math-container\">$n$</span> is a solution for <span class=\"math-container\">$k=k_a$</span>. And if <span class=\"math-container\">$n$</span> is a solution for <span class=\"math-container\">$k=k_a+1$</span> with <span class=\"math-container\">$n&lt;2^{2k_a+1}$</span> and <span class=\"math-container\">$n$</span> even, then <span class=\"math-container\">$n/2$</span> is a solution for <span class=\"math-container\">$k=k_a$</span>. That explains why more often than not, if <span class=\"math-container\">$n=n_a$</span> is a solution for <span class=\"math-container\">$k=k_a$</span>, then <span class=\"math-container\">$n=n_a$</span> or/and <span class=\"math-container\">$n=2n_a$</span> is/are a solution for <span class=\"math-container\">$k=k_a+1$</span>.</p>\n" }, { "answer_id": 259677, "score": 6, "text": "<p>Not an answer, but too long for a comment.</p>\n\n<p>I can't see any tricks to do this other than a brute force computational approach. The naive approach would be to loop from $2^{96}$ to $2^{128}$ squaring each integer and checking. A more efficient approach in practice is to start with the number $2^{192}-2^{64}$ and then loop up to $2^{256}-2^{64}$ in steps of $2^{192}$ -- this only involves a loop of length $2^{64}$ rather than one of length $2^{128}$. For each term $n$ in the loop (thought of as a good approximation to a square) one computes the smallest integer square which is at least $n$ and then checks to see if the difference is less than $2^{64}$. But a loop of length $2^{64}$ is too long for me.</p>\n\n<p>Here's some pari-gp code which takes an even length $n=2m$ and solves the analogous problem of searching for $n$-bit numbers which have 1's from the bit controlling $2^m$ (the \"$m+1$st bit, I think, in the language of the OP) to the bit controlling $2^{3m-1}$.</p>\n\n<pre><code>g(n)=m=n/2;d=2^(3*m)-2^m;r=0;forstep(i0=d,2^(4*m)-2^m,2^(3*m),if(issquare(i0),s=sqrtint(i0),s=sqrtint(i0)+1);if(s^2-i0&lt;2^m,print(\"yes -- s=\",s);r=1;break()));if(r==0,print(\"no\"));r\n</code></pre>\n\n<p>It (rather inelegantly) prints the smallest number it finds, if it finds one, and returns 1 or 0 depending on whether it finds one or not. It runs to $n=60$ quickly.</p>\n\n<p>What motivated me to post this as an answer was the observation that in fact for most values of $n$ it could find an $n$-bit number which did the job. There's nothing for 2,4,6 but 181 works for 8 and 10 (this is related to the famous fact that $181^2=2^{15}-7$). There are then failures for $n=12,14,16,18$ but after that the garden gets rosier -- the numbers 699733 (twice),1853638,707276,170076312 work for 20,22,24,26,28, we have another failure at 30, and then 2863355221 works for 32,34; another failure at 36, success at 38, failure at 40 and then another big string of successes -- everything from 42 to 62 works. For example 363103060890424251 works for 60 and 62; in binary the square of this is 110010110010001100111000111111111111111111111111111111111111111111111111111111111111111001111101101110100101010011001 . This somehow indicates the problem in general -- you have a bunch of 1's in the middle but things look pretty random outside that region. The loop I posted above brute forces the first string of digits and then looks for the nearest square (because there will be at most one string of digits at the end making the number square). </p>\n\n<p>I do not know a more sensible way of approaching this problem, however given that for most small values of $n$ there did seem to be a solution one might be cautiously optimistic that such an integer could exist. I suppose that on average the moment one passes $2^{96}$ the middle digits are kind-of random, so one might ask instead if you have an event which has a $2^{-128}$ chance of happening but you have $2^{128}-2^{96}$ attempts at it, what are the chances of you being lucky, and they're about $1-1/e$. So a sporting chance!</p>\n" }, { "answer_id": 259822, "score": 6, "text": "<p>I'd like to point out why the case of middle bits being all ones is somewhat special and different from other fixed values of them. Also, there is an approach for finding a suitable numbers that may not exist for other fixed values of middle bits.</p>\n\n<p>I'll use the notation of Kevin and additionally require that $m$ is even, say, $m=2k$ (in OP's question we have $m=64$ and $k=32$).</p>\n\n<p>We need to find integers $x,y$ such that $0\\leq x,y&lt;2^m$ and \n$$x\\cdot 2^{3m} + (2^{3m}-2^m) + y=z^2$$\nfor some integer $z$. In other words, we have\n$$0&lt;(x+1) - \\frac{z^2}{2^{3m}} = \\frac{(2^m-y)}{2^{3m}} &lt; \\frac{1}{2^{2m}}.$$\nFactoring the l.h.s. as $(\\sqrt{x+1}-\\frac{z}{2^{3k}})(\\sqrt{x+1}+\\frac{z}{2^{3k}})$ and noticing that the latter factor is at least 2 (for $x&gt;0$), we get\n$$0&lt;\\sqrt{x+1} - \\frac{z}{2^{3k}} &lt; \\frac{1}{2^{4k+1}}.$$\nThis tells us that $\\frac{z}{2^{3k}}$ is a very good rational approximation to $\\sqrt{x+1}$. </p>\n\n<p>The above analysis may suggest to search for $\\frac{z}{2^{3k}}$ among convergents and semiconvergents to a square root of an integer. We can base this search on two facts:</p>\n\n<ul>\n<li>A continued fraction for a square root has special forms: $[a;\\overline{2a}]$, $[a;\\overline{b,2a}]$, $[a;\\overline{b,b,2a}]$, $[a;\\overline{b,c,b,2a}]$, $[a;\\overline{b,c,c,b,2a}]$, etc.</li>\n<li>Denominators of (semi)convergents satisfy a simple recurrence relation (involving terms of the continued fraction). </li>\n</ul>\n\n<p>So, for each of the above continued fractions, we may to try find values of $a,b,c,\\dots$ such that $2^t$ with $t\\leq 3k$ appears among the denominators of (semi)convergents, from which will further get $z$ (as the numerator times $2^{3k-t}$) and hopefully solve the problem.</p>\n\n<p><strong>Remarks.</strong> </p>\n\n<ol>\n<li>Values $t\\leq 2k$ may work only for convergents, not semiconvergents (the rational approximation in this case becomes so good that only convergents may satisfy it).</li>\n<li>There is a number of underwater stones here such as (i) not every set of values $a,b,c,\\dots$ guarantee that we have a a continued fraction of the square root of an integer (in general, it's the square root of a rational); (ii) semiconvergents may not guarantee that the approximation is well enough for our purposes; etc. </li>\n</ol>\n" }, { "answer_id": 418082, "score": 7, "text": "<blockquote>\n<p>Is there a positive 128-bit integer whose square has all middle bits equal to 1?</p>\n</blockquote>\n<p><em><strong>YES</strong></em>. One is AAAAAAAAAAAAAAAB5555555555555555<sub>16</sub>, which square is 71C71C71C71C71C7FFFFFFFFFFFFFFFFFFFFFFFFFFFFFFFF8E38E38E38E38E39<sub>16</sub>.</p>\n<hr />\n<p>It's asked if there are integers <span class=\"math-container\">$n$</span> such that <span class=\"math-container\">$0&lt;n&lt;2^{2k}$</span> and <span class=\"math-container\">$(n^2\\bmod2^{3k})\\ge2^{3k}-2^k$</span>, in the case <span class=\"math-container\">$k=128/2$</span>.</p>\n<p>As noted in <a href=\"https://mathoverflow.net/posts/comments/1073379?noredirect=1\">comment by Max Alekseyev</a>, that holds for <span class=\"math-container\">$k\\equiv 4\\pmod6$</span> (thus including the question's <span class=\"math-container\">$k=128/2$</span>), because the integer\n<span class=\"math-container\">$$n=\\frac{2^{2k+1}+2^{k+1}-1}3$$</span>\ndoes the trick. It's representation in base <span class=\"math-container\">$4$</span> consists of <span class=\"math-container\">$k/2-1$</span> digits <span class=\"math-container\">$2$</span>, one digit <span class=\"math-container\">$3$</span>, and <span class=\"math-container\">$k/2$</span> digits <span class=\"math-container\">$1$</span>. It holds\n<span class=\"math-container\">$$n^2=\\frac{8^k-1}{9}2^{k+2}+\\frac{8^{k+1}+1}9$$</span>\nwhich representation in base <span class=\"math-container\">$4$</span> consists of the digit <span class=\"math-container\">$1$</span>; the digits <span class=\"math-container\">$301$</span> repeated <span class=\"math-container\">$(k-4)/6$</span> times; the digit <span class=\"math-container\">$3$</span> repeated <span class=\"math-container\">$1+k$</span> times; the digit <span class=\"math-container\">$2$</span>; the digits <span class=\"math-container\">$032$</span> repeated <span class=\"math-container\">$(k-4)/6$</span> times; the digit <span class=\"math-container\">$1$</span>. It follows the center <span class=\"math-container\">$2k$</span> bits are set, with three more on the left and one more on the right.</p>\n<p>We can't use this to conclude exactly how many <span class=\"math-container\">$n$</span> there are, or/and for other values of <span class=\"math-container\">$k$</span>, or/and for an arbitrary value in the middle of the square.</p>\n<p>Here is a brute force search for <span class=\"math-container\">$k\\le21$</span>:</p>\n<div class=\"s-table-container\">\n<table class=\"s-table\">\n<thead>\n<tr>\n<th style=\"text-align: center;\"><span class=\"math-container\">$k$</span> with a solution</th>\n<th style=\"text-align: center;\">#solution(s)</th>\n<th style=\"text-align: right;\">possible <span class=\"math-container\">$n$</span></th>\n</tr>\n</thead>\n<tbody>\n<tr>\n<td style=\"text-align: center;\">4</td>\n<td style=\"text-align: center;\">1</td>\n<td style=\"text-align: right;\">10110101<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">5</td>\n<td style=\"text-align: center;\">2</td>\n<td style=\"text-align: right;\">0010110101<sub>2</sub><br>0101101010<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">10</td>\n<td style=\"text-align: center;\">2</td>\n<td style=\"text-align: right;\">10101010110101010101<sub>2</sub><br>11100010010001100011<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">11</td>\n<td style=\"text-align: center;\">4</td>\n<td style=\"text-align: right;\">0010101010110101010101<sub>2</sub><br>0011100010010001100011<sub>2</sub><br>0101010101101010101010<sub>2</sub><br>0111000100100011000110<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">12</td>\n<td style=\"text-align: center;\">3</td>\n<td style=\"text-align: right;\">000111000100100011000110<sub>2</sub><br>010010001110001000101011<sub>2</sub><br>110001101110000001111001<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">13</td>\n<td style=\"text-align: center;\">3</td>\n<td style=\"text-align: right;\">00001110001001000110001100<sub>2</sub><br>00110001101110000001111001<sub>2</sub><br>01100011011100000011110010<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">14</td>\n<td style=\"text-align: center;\">1</td>\n<td style=\"text-align: right;\">1010001000110010100010011000<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">16</td>\n<td style=\"text-align: center;\">1</td>\n<td style=\"text-align: right;\">10101010101010110101010101010101<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">17</td>\n<td style=\"text-align: center;\">2</td>\n<td style=\"text-align: right;\">0010101010101010110101010101010101<sub>2</sub><br>0101010101010101101010101010101010<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">19</td>\n<td style=\"text-align: center;\">1</td>\n<td style=\"text-align: right;\">11110010011001000101100100110110000011<sub>2</sub></td>\n</tr>\n<tr>\n<td style=\"text-align: center;\">21</td>\n<td style=\"text-align: center;\">4</td>\n<td style=\"text-align: right;\">010100000101011011101010011000101011000111<sub>2</sub><br>011101100111011110011000000000010111000101<sub>2</sub><br>101000001010110111010100110001010110001110<sub>2</sub><br>111101110010111001100001001101110110100001<sub>2</sub></td>\n</tr>\n</tbody>\n</table>\n</div>\n<p>We see and could prove that if <span class=\"math-container\">$n$</span> is a solution for <span class=\"math-container\">$k=k_a+1$</span> with <span class=\"math-container\">$n&lt;2^{2k_a}$</span>, then <span class=\"math-container\">$n$</span> is a solution for <span class=\"math-container\">$k=k_a$</span>. And if <span class=\"math-container\">$n$</span> is a solution for <span class=\"math-container\">$k=k_a+1$</span> with <span class=\"math-container\">$n&lt;2^{2k_a+1}$</span> and <span class=\"math-container\">$n$</span> even, then <span class=\"math-container\">$n/2$</span> is a solution for <span class=\"math-container\">$k=k_a$</span>. That explains why more often than not, if <span class=\"math-container\">$n=n_a$</span> is a solution for <span class=\"math-container\">$k=k_a$</span>, then <span class=\"math-container\">$n=n_a$</span> or/and <span class=\"math-container\">$n=2n_a$</span> is/are a solution for <span class=\"math-container\">$k=k_a+1$</span>.</p>\n" }, { "answer_id": 259677, "score": 6, "text": "<p>Not an answer, but too long for a comment.</p>\n\n<p>I can't see any tricks to do this other than a brute force computational approach. The naive approach would be to loop from $2^{96}$ to $2^{128}$ squaring each integer and checking. A more efficient approach in practice is to start with the number $2^{192}-2^{64}$ and then loop up to $2^{256}-2^{64}$ in steps of $2^{192}$ -- this only involves a loop of length $2^{64}$ rather than one of length $2^{128}$. For each term $n$ in the loop (thought of as a good approximation to a square) one computes the smallest integer square which is at least $n$ and then checks to see if the difference is less than $2^{64}$. But a loop of length $2^{64}$ is too long for me.</p>\n\n<p>Here's some pari-gp code which takes an even length $n=2m$ and solves the analogous problem of searching for $n$-bit numbers which have 1's from the bit controlling $2^m$ (the \"$m+1$st bit, I think, in the language of the OP) to the bit controlling $2^{3m-1}$.</p>\n\n<pre><code>g(n)=m=n/2;d=2^(3*m)-2^m;r=0;forstep(i0=d,2^(4*m)-2^m,2^(3*m),if(issquare(i0),s=sqrtint(i0),s=sqrtint(i0)+1);if(s^2-i0&lt;2^m,print(\"yes -- s=\",s);r=1;break()));if(r==0,print(\"no\"));r\n</code></pre>\n\n<p>It (rather inelegantly) prints the smallest number it finds, if it finds one, and returns 1 or 0 depending on whether it finds one or not. It runs to $n=60$ quickly.</p>\n\n<p>What motivated me to post this as an answer was the observation that in fact for most values of $n$ it could find an $n$-bit number which did the job. There's nothing for 2,4,6 but 181 works for 8 and 10 (this is related to the famous fact that $181^2=2^{15}-7$). There are then failures for $n=12,14,16,18$ but after that the garden gets rosier -- the numbers 699733 (twice),1853638,707276,170076312 work for 20,22,24,26,28, we have another failure at 30, and then 2863355221 works for 32,34; another failure at 36, success at 38, failure at 40 and then another big string of successes -- everything from 42 to 62 works. For example 363103060890424251 works for 60 and 62; in binary the square of this is 110010110010001100111000111111111111111111111111111111111111111111111111111111111111111001111101101110100101010011001 . This somehow indicates the problem in general -- you have a bunch of 1's in the middle but things look pretty random outside that region. The loop I posted above brute forces the first string of digits and then looks for the nearest square (because there will be at most one string of digits at the end making the number square). </p>\n\n<p>I do not know a more sensible way of approaching this problem, however given that for most small values of $n$ there did seem to be a solution one might be cautiously optimistic that such an integer could exist. I suppose that on average the moment one passes $2^{96}$ the middle digits are kind-of random, so one might ask instead if you have an event which has a $2^{-128}$ chance of happening but you have $2^{128}-2^{96}$ attempts at it, what are the chances of you being lucky, and they're about $1-1/e$. So a sporting chance!</p>\n" }, { "answer_id": 259822, "score": 6, "text": "<p>I'd like to point out why the case of middle bits being all ones is somewhat special and different from other fixed values of them. Also, there is an approach for finding a suitable numbers that may not exist for other fixed values of middle bits.</p>\n\n<p>I'll use the notation of Kevin and additionally require that $m$ is even, say, $m=2k$ (in OP's question we have $m=64$ and $k=32$).</p>\n\n<p>We need to find integers $x,y$ such that $0\\leq x,y&lt;2^m$ and \n$$x\\cdot 2^{3m} + (2^{3m}-2^m) + y=z^2$$\nfor some integer $z$. In other words, we have\n$$0&lt;(x+1) - \\frac{z^2}{2^{3m}} = \\frac{(2^m-y)}{2^{3m}} &lt; \\frac{1}{2^{2m}}.$$\nFactoring the l.h.s. as $(\\sqrt{x+1}-\\frac{z}{2^{3k}})(\\sqrt{x+1}+\\frac{z}{2^{3k}})$ and noticing that the latter factor is at least 2 (for $x&gt;0$), we get\n$$0&lt;\\sqrt{x+1} - \\frac{z}{2^{3k}} &lt; \\frac{1}{2^{4k+1}}.$$\nThis tells us that $\\frac{z}{2^{3k}}$ is a very good rational approximation to $\\sqrt{x+1}$. </p>\n\n<p>The above analysis may suggest to search for $\\frac{z}{2^{3k}}$ among convergents and semiconvergents to a square root of an integer. We can base this search on two facts:</p>\n\n<ul>\n<li>A continued fraction for a square root has special forms: $[a;\\overline{2a}]$, $[a;\\overline{b,2a}]$, $[a;\\overline{b,b,2a}]$, $[a;\\overline{b,c,b,2a}]$, $[a;\\overline{b,c,c,b,2a}]$, etc.</li>\n<li>Denominators of (semi)convergents satisfy a simple recurrence relation (involving terms of the continued fraction). </li>\n</ul>\n\n<p>So, for each of the above continued fractions, we may to try find values of $a,b,c,\\dots$ such that $2^t$ with $t\\leq 3k$ appears among the denominators of (semi)convergents, from which will further get $z$ (as the numerator times $2^{3k-t}$) and hopefully solve the problem.</p>\n\n<p><strong>Remarks.</strong> </p>\n\n<ol>\n<li>Values $t\\leq 2k$ may work only for convergents, not semiconvergents (the rational approximation in this case becomes so good that only convergents may satisfy it).</li>\n<li>There is a number of underwater stones here such as (i) not every set of values $a,b,c,\\dots$ guarantee that we have a a continued fraction of the square root of an integer (in general, it's the square root of a rational); (ii) semiconvergents may not guarantee that the approximation is well enough for our purposes; etc. </li>\n</ol>\n" } ]
Question: Mid-Square with all bits set Answer 1: <blockquote> <p>Is there a positive 128-bit integer whose square has all middle bits equal to 1?</p> </blockquote> <p><em><strong>YES</strong></em>. One is AAAAAAAAAAAAAAAB5555555555555555<sub>16</sub>, which square is 71C71C71C71C71C7FFFFFFFFFFFFFFFFFFFFFFFFFFFFFFFF8E38E38E38E38E39<sub>16</sub>.</p> <hr /> <p>It's asked if there are integers <span class="math-container">$n$</span> such that <span class="math-container">$0&lt;n&lt;2^{2k}$</span> and <span class="math-container">$(n^2\bmod2^{3k})\ge2^{3k}-2^k$</span>, in the case <span class="math-container">$k=128/2$</span>.</p> <p>As noted in <a href="https://mathoverflow.net/posts/comments/1073379?noredirect=1">comment by Max Alekseyev</a>, that holds for <span class="math-container">$k\equiv 4\pmod6$</span> (thus including the question's <span class="math-container">$k=128/2$</span>), because the integer <span class="math-container">$$n=\frac{2^{2k+1}+2^{k+1}-1}3$$</span> does the trick. It's representation in base <span class="math-container">$4$</span> consists of <span class="math-container">$k/2-1$</span> digits <span class="math-container">$2$</span>, one digit <span class="math-container">$3$</span>, and <span class="math-container">$k/2$</span> digits <span class="math-container">$1$</span>. It holds <span class="math-container">$$n^2=\frac{8^k-1}{9}2^{k+2}+\frac{8^{k+1}+1}9$$</span> which representation in base <span class="math-container">$4$</span> consists of the digit <span class="math-container">$1$</span>; the digits <span class="math-container">$301$</span> repeated <span class="math-container">$(k-4)/6$</span> times; the digit <span class="math-container">$3$</span> repeated <span class="math-container">$1+k$</span> times; the digit <span class="math-container">$2$</span>; the digits <span class="math-container">$032$</span> repeated <span class="math-container">$(k-4)/6$</span> times; the digit <span class="math-container">$1$</span>. It follows the center <span class="math-container">$2k$</span> bits are set, with three more on the left and one more on the right.</p> <p>We can't use this to conclude exactly how many <span class="math-container">$n$</span> there are, or/and for other values of <span class="math-container">$k$</span>, or/and for an arbitrary value in the middle of the square.</p> <p>Here is a brute force search for <span class="math-container">$k\le21$</span>:</p> <div class="s-table-container"> <table class="s-table"> <thead> <tr> <th style="text-align: center;"><span class="math-container">$k$</span> with a solution</th> <th style="text-align: center;">#solution(s)</th> <th style="text-align: right;">possible <span class="math-container">$n$</span></th> </tr> </thead> <tbody> <tr> <td style="text-align: center;">4</td> <td style="text-align: center;">1</td> <td style="text-align: right;">10110101<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">5</td> <td style="text-align: center;">2</td> <td style="text-align: right;">0010110101<sub>2</sub><br>0101101010<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">10</td> <td style="text-align: center;">2</td> <td style="text-align: right;">10101010110101010101<sub>2</sub><br>11100010010001100011<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">11</td> <td style="text-align: center;">4</td> <td style="text-align: right;">0010101010110101010101<sub>2</sub><br>0011100010010001100011<sub>2</sub><br>0101010101101010101010<sub>2</sub><br>0111000100100011000110<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">12</td> <td style="text-align: center;">3</td> <td style="text-align: right;">000111000100100011000110<sub>2</sub><br>010010001110001000101011<sub>2</sub><br>110001101110000001111001<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">13</td> <td style="text-align: center;">3</td> <td style="text-align: right;">00001110001001000110001100<sub>2</sub><br>00110001101110000001111001<sub>2</sub><br>01100011011100000011110010<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">14</td> <td style="text-align: center;">1</td> <td style="text-align: right;">1010001000110010100010011000<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">16</td> <td style="text-align: center;">1</td> <td style="text-align: right;">10101010101010110101010101010101<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">17</td> <td style="text-align: center;">2</td> <td style="text-align: right;">0010101010101010110101010101010101<sub>2</sub><br>0101010101010101101010101010101010<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">19</td> <td style="text-align: center;">1</td> <td style="text-align: right;">11110010011001000101100100110110000011<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">21</td> <td style="text-align: center;">4</td> <td style="text-align: right;">010100000101011011101010011000101011000111<sub>2</sub><br>011101100111011110011000000000010111000101<sub>2</sub><br>101000001010110111010100110001010110001110<sub>2</sub><br>111101110010111001100001001101110110100001<sub>2</sub></td> </tr> </tbody> </table> </div> <p>We see and could prove that if <span class="math-container">$n$</span> is a solution for <span class="math-container">$k=k_a+1$</span> with <span class="math-container">$n&lt;2^{2k_a}$</span>, then <span class="math-container">$n$</span> is a solution for <span class="math-container">$k=k_a$</span>. And if <span class="math-container">$n$</span> is a solution for <span class="math-container">$k=k_a+1$</span> with <span class="math-container">$n&lt;2^{2k_a+1}$</span> and <span class="math-container">$n$</span> even, then <span class="math-container">$n/2$</span> is a solution for <span class="math-container">$k=k_a$</span>. That explains why more often than not, if <span class="math-container">$n=n_a$</span> is a solution for <span class="math-container">$k=k_a$</span>, then <span class="math-container">$n=n_a$</span> or/and <span class="math-container">$n=2n_a$</span> is/are a solution for <span class="math-container">$k=k_a+1$</span>.</p> Answer 2: <p>Not an answer, but too long for a comment.</p> <p>I can't see any tricks to do this other than a brute force computational approach. The naive approach would be to loop from $2^{96}$ to $2^{128}$ squaring each integer and checking. A more efficient approach in practice is to start with the number $2^{192}-2^{64}$ and then loop up to $2^{256}-2^{64}$ in steps of $2^{192}$ -- this only involves a loop of length $2^{64}$ rather than one of length $2^{128}$. For each term $n$ in the loop (thought of as a good approximation to a square) one computes the smallest integer square which is at least $n$ and then checks to see if the difference is less than $2^{64}$. But a loop of length $2^{64}$ is too long for me.</p> <p>Here's some pari-gp code which takes an even length $n=2m$ and solves the analogous problem of searching for $n$-bit numbers which have 1's from the bit controlling $2^m$ (the "$m+1$st bit, I think, in the language of the OP) to the bit controlling $2^{3m-1}$.</p> <pre><code>g(n)=m=n/2;d=2^(3*m)-2^m;r=0;forstep(i0=d,2^(4*m)-2^m,2^(3*m),if(issquare(i0),s=sqrtint(i0),s=sqrtint(i0)+1);if(s^2-i0&lt;2^m,print("yes -- s=",s);r=1;break()));if(r==0,print("no"));r </code></pre> <p>It (rather inelegantly) prints the smallest number it finds, if it finds one, and returns 1 or 0 depending on whether it finds one or not. It runs to $n=60$ quickly.</p> <p>What motivated me to post this as an answer was the observation that in fact for most values of $n$ it could find an $n$-bit number which did the job. There's nothing for 2,4,6 but 181 works for 8 and 10 (this is related to the famous fact that $181^2=2^{15}-7$). There are then failures for $n=12,14,16,18$ but after that the garden gets rosier -- the numbers 699733 (twice),1853638,707276,170076312 work for 20,22,24,26,28, we have another failure at 30, and then 2863355221 works for 32,34; another failure at 36, success at 38, failure at 40 and then another big string of successes -- everything from 42 to 62 works. For example 363103060890424251 works for 60 and 62; in binary the square of this is 110010110010001100111000111111111111111111111111111111111111111111111111111111111111111001111101101110100101010011001 . This somehow indicates the problem in general -- you have a bunch of 1's in the middle but things look pretty random outside that region. The loop I posted above brute forces the first string of digits and then looks for the nearest square (because there will be at most one string of digits at the end making the number square). </p> <p>I do not know a more sensible way of approaching this problem, however given that for most small values of $n$ there did seem to be a solution one might be cautiously optimistic that such an integer could exist. I suppose that on average the moment one passes $2^{96}$ the middle digits are kind-of random, so one might ask instead if you have an event which has a $2^{-128}$ chance of happening but you have $2^{128}-2^{96}$ attempts at it, what are the chances of you being lucky, and they're about $1-1/e$. So a sporting chance!</p> Answer 3: <p>I'd like to point out why the case of middle bits being all ones is somewhat special and different from other fixed values of them. Also, there is an approach for finding a suitable numbers that may not exist for other fixed values of middle bits.</p> <p>I'll use the notation of Kevin and additionally require that $m$ is even, say, $m=2k$ (in OP's question we have $m=64$ and $k=32$).</p> <p>We need to find integers $x,y$ such that $0\leq x,y&lt;2^m$ and $$x\cdot 2^{3m} + (2^{3m}-2^m) + y=z^2$$ for some integer $z$. In other words, we have $$0&lt;(x+1) - \frac{z^2}{2^{3m}} = \frac{(2^m-y)}{2^{3m}} &lt; \frac{1}{2^{2m}}.$$ Factoring the l.h.s. as $(\sqrt{x+1}-\frac{z}{2^{3k}})(\sqrt{x+1}+\frac{z}{2^{3k}})$ and noticing that the latter factor is at least 2 (for $x&gt;0$), we get $$0&lt;\sqrt{x+1} - \frac{z}{2^{3k}} &lt; \frac{1}{2^{4k+1}}.$$ This tells us that $\frac{z}{2^{3k}}$ is a very good rational approximation to $\sqrt{x+1}$. </p> <p>The above analysis may suggest to search for $\frac{z}{2^{3k}}$ among convergents and semiconvergents to a square root of an integer. We can base this search on two facts:</p> <ul> <li>A continued fraction for a square root has special forms: $[a;\overline{2a}]$, $[a;\overline{b,2a}]$, $[a;\overline{b,b,2a}]$, $[a;\overline{b,c,b,2a}]$, $[a;\overline{b,c,c,b,2a}]$, etc.</li> <li>Denominators of (semi)convergents satisfy a simple recurrence relation (involving terms of the continued fraction). </li> </ul> <p>So, for each of the above continued fractions, we may to try find values of $a,b,c,\dots$ such that $2^t$ with $t\leq 3k$ appears among the denominators of (semi)convergents, from which will further get $z$ (as the numerator times $2^{3k-t}$) and hopefully solve the problem.</p> <p><strong>Remarks.</strong> </p> <ol> <li>Values $t\leq 2k$ may work only for convergents, not semiconvergents (the rational approximation in this case becomes so good that only convergents may satisfy it).</li> <li>There is a number of underwater stones here such as (i) not every set of values $a,b,c,\dots$ guarantee that we have a a continued fraction of the square root of an integer (in general, it's the square root of a rational); (ii) semiconvergents may not guarantee that the approximation is well enough for our purposes; etc. </li> </ol> Answer 4: <blockquote> <p>Is there a positive 128-bit integer whose square has all middle bits equal to 1?</p> </blockquote> <p><em><strong>YES</strong></em>. One is AAAAAAAAAAAAAAAB5555555555555555<sub>16</sub>, which square is 71C71C71C71C71C7FFFFFFFFFFFFFFFFFFFFFFFFFFFFFFFF8E38E38E38E38E39<sub>16</sub>.</p> <hr /> <p>It's asked if there are integers <span class="math-container">$n$</span> such that <span class="math-container">$0&lt;n&lt;2^{2k}$</span> and <span class="math-container">$(n^2\bmod2^{3k})\ge2^{3k}-2^k$</span>, in the case <span class="math-container">$k=128/2$</span>.</p> <p>As noted in <a href="https://mathoverflow.net/posts/comments/1073379?noredirect=1">comment by Max Alekseyev</a>, that holds for <span class="math-container">$k\equiv 4\pmod6$</span> (thus including the question's <span class="math-container">$k=128/2$</span>), because the integer <span class="math-container">$$n=\frac{2^{2k+1}+2^{k+1}-1}3$$</span> does the trick. It's representation in base <span class="math-container">$4$</span> consists of <span class="math-container">$k/2-1$</span> digits <span class="math-container">$2$</span>, one digit <span class="math-container">$3$</span>, and <span class="math-container">$k/2$</span> digits <span class="math-container">$1$</span>. It holds <span class="math-container">$$n^2=\frac{8^k-1}{9}2^{k+2}+\frac{8^{k+1}+1}9$$</span> which representation in base <span class="math-container">$4$</span> consists of the digit <span class="math-container">$1$</span>; the digits <span class="math-container">$301$</span> repeated <span class="math-container">$(k-4)/6$</span> times; the digit <span class="math-container">$3$</span> repeated <span class="math-container">$1+k$</span> times; the digit <span class="math-container">$2$</span>; the digits <span class="math-container">$032$</span> repeated <span class="math-container">$(k-4)/6$</span> times; the digit <span class="math-container">$1$</span>. It follows the center <span class="math-container">$2k$</span> bits are set, with three more on the left and one more on the right.</p> <p>We can't use this to conclude exactly how many <span class="math-container">$n$</span> there are, or/and for other values of <span class="math-container">$k$</span>, or/and for an arbitrary value in the middle of the square.</p> <p>Here is a brute force search for <span class="math-container">$k\le21$</span>:</p> <div class="s-table-container"> <table class="s-table"> <thead> <tr> <th style="text-align: center;"><span class="math-container">$k$</span> with a solution</th> <th style="text-align: center;">#solution(s)</th> <th style="text-align: right;">possible <span class="math-container">$n$</span></th> </tr> </thead> <tbody> <tr> <td style="text-align: center;">4</td> <td style="text-align: center;">1</td> <td style="text-align: right;">10110101<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">5</td> <td style="text-align: center;">2</td> <td style="text-align: right;">0010110101<sub>2</sub><br>0101101010<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">10</td> <td style="text-align: center;">2</td> <td style="text-align: right;">10101010110101010101<sub>2</sub><br>11100010010001100011<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">11</td> <td style="text-align: center;">4</td> <td style="text-align: right;">0010101010110101010101<sub>2</sub><br>0011100010010001100011<sub>2</sub><br>0101010101101010101010<sub>2</sub><br>0111000100100011000110<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">12</td> <td style="text-align: center;">3</td> <td style="text-align: right;">000111000100100011000110<sub>2</sub><br>010010001110001000101011<sub>2</sub><br>110001101110000001111001<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">13</td> <td style="text-align: center;">3</td> <td style="text-align: right;">00001110001001000110001100<sub>2</sub><br>00110001101110000001111001<sub>2</sub><br>01100011011100000011110010<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">14</td> <td style="text-align: center;">1</td> <td style="text-align: right;">1010001000110010100010011000<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">16</td> <td style="text-align: center;">1</td> <td style="text-align: right;">10101010101010110101010101010101<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">17</td> <td style="text-align: center;">2</td> <td style="text-align: right;">0010101010101010110101010101010101<sub>2</sub><br>0101010101010101101010101010101010<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">19</td> <td style="text-align: center;">1</td> <td style="text-align: right;">11110010011001000101100100110110000011<sub>2</sub></td> </tr> <tr> <td style="text-align: center;">21</td> <td style="text-align: center;">4</td> <td style="text-align: right;">010100000101011011101010011000101011000111<sub>2</sub><br>011101100111011110011000000000010111000101<sub>2</sub><br>101000001010110111010100110001010110001110<sub>2</sub><br>111101110010111001100001001101110110100001<sub>2</sub></td> </tr> </tbody> </table> </div> <p>We see and could prove that if <span class="math-container">$n$</span> is a solution for <span class="math-container">$k=k_a+1$</span> with <span class="math-container">$n&lt;2^{2k_a}$</span>, then <span class="math-container">$n$</span> is a solution for <span class="math-container">$k=k_a$</span>. And if <span class="math-container">$n$</span> is a solution for <span class="math-container">$k=k_a+1$</span> with <span class="math-container">$n&lt;2^{2k_a+1}$</span> and <span class="math-container">$n$</span> even, then <span class="math-container">$n/2$</span> is a solution for <span class="math-container">$k=k_a$</span>. That explains why more often than not, if <span class="math-container">$n=n_a$</span> is a solution for <span class="math-container">$k=k_a$</span>, then <span class="math-container">$n=n_a$</span> or/and <span class="math-container">$n=2n_a$</span> is/are a solution for <span class="math-container">$k=k_a+1$</span>.</p> Answer 5: <p>Not an answer, but too long for a comment.</p> <p>I can't see any tricks to do this other than a brute force computational approach. The naive approach would be to loop from $2^{96}$ to $2^{128}$ squaring each integer and checking. A more efficient approach in practice is to start with the number $2^{192}-2^{64}$ and then loop up to $2^{256}-2^{64}$ in steps of $2^{192}$ -- this only involves a loop of length $2^{64}$ rather than one of length $2^{128}$. For each term $n$ in the loop (thought of as a good approximation to a square) one computes the smallest integer square which is at least $n$ and then checks to see if the difference is less than $2^{64}$. But a loop of length $2^{64}$ is too long for me.</p> <p>Here's some pari-gp code which takes an even length $n=2m$ and solves the analogous problem of searching for $n$-bit numbers which have 1's from the bit controlling $2^m$ (the "$m+1$st bit, I think, in the language of the OP) to the bit controlling $2^{3m-1}$.</p> <pre><code>g(n)=m=n/2;d=2^(3*m)-2^m;r=0;forstep(i0=d,2^(4*m)-2^m,2^(3*m),if(issquare(i0),s=sqrtint(i0),s=sqrtint(i0)+1);if(s^2-i0&lt;2^m,print("yes -- s=",s);r=1;break()));if(r==0,print("no"));r </code></pre> <p>It (rather inelegantly) prints the smallest number it finds, if it finds one, and returns 1 or 0 depending on whether it finds one or not. It runs to $n=60$ quickly.</p> <p>What motivated me to post this as an answer was the observation that in fact for most values of $n$ it could find an $n$-bit number which did the job. There's nothing for 2,4,6 but 181 works for 8 and 10 (this is related to the famous fact that $181^2=2^{15}-7$). There are then failures for $n=12,14,16,18$ but after that the garden gets rosier -- the numbers 699733 (twice),1853638,707276,170076312 work for 20,22,24,26,28, we have another failure at 30, and then 2863355221 works for 32,34; another failure at 36, success at 38, failure at 40 and then another big string of successes -- everything from 42 to 62 works. For example 363103060890424251 works for 60 and 62; in binary the square of this is 110010110010001100111000111111111111111111111111111111111111111111111111111111111111111001111101101110100101010011001 . This somehow indicates the problem in general -- you have a bunch of 1's in the middle but things look pretty random outside that region. The loop I posted above brute forces the first string of digits and then looks for the nearest square (because there will be at most one string of digits at the end making the number square). </p> <p>I do not know a more sensible way of approaching this problem, however given that for most small values of $n$ there did seem to be a solution one might be cautiously optimistic that such an integer could exist. I suppose that on average the moment one passes $2^{96}$ the middle digits are kind-of random, so one might ask instead if you have an event which has a $2^{-128}$ chance of happening but you have $2^{128}-2^{96}$ attempts at it, what are the chances of you being lucky, and they're about $1-1/e$. So a sporting chance!</p> Answer 6: <p>I'd like to point out why the case of middle bits being all ones is somewhat special and different from other fixed values of them. Also, there is an approach for finding a suitable numbers that may not exist for other fixed values of middle bits.</p> <p>I'll use the notation of Kevin and additionally require that $m$ is even, say, $m=2k$ (in OP's question we have $m=64$ and $k=32$).</p> <p>We need to find integers $x,y$ such that $0\leq x,y&lt;2^m$ and $$x\cdot 2^{3m} + (2^{3m}-2^m) + y=z^2$$ for some integer $z$. In other words, we have $$0&lt;(x+1) - \frac{z^2}{2^{3m}} = \frac{(2^m-y)}{2^{3m}} &lt; \frac{1}{2^{2m}}.$$ Factoring the l.h.s. as $(\sqrt{x+1}-\frac{z}{2^{3k}})(\sqrt{x+1}+\frac{z}{2^{3k}})$ and noticing that the latter factor is at least 2 (for $x&gt;0$), we get $$0&lt;\sqrt{x+1} - \frac{z}{2^{3k}} &lt; \frac{1}{2^{4k+1}}.$$ This tells us that $\frac{z}{2^{3k}}$ is a very good rational approximation to $\sqrt{x+1}$. </p> <p>The above analysis may suggest to search for $\frac{z}{2^{3k}}$ among convergents and semiconvergents to a square root of an integer. We can base this search on two facts:</p> <ul> <li>A continued fraction for a square root has special forms: $[a;\overline{2a}]$, $[a;\overline{b,2a}]$, $[a;\overline{b,b,2a}]$, $[a;\overline{b,c,b,2a}]$, $[a;\overline{b,c,c,b,2a}]$, etc.</li> <li>Denominators of (semi)convergents satisfy a simple recurrence relation (involving terms of the continued fraction). </li> </ul> <p>So, for each of the above continued fractions, we may to try find values of $a,b,c,\dots$ such that $2^t$ with $t\leq 3k$ appears among the denominators of (semi)convergents, from which will further get $z$ (as the numerator times $2^{3k-t}$) and hopefully solve the problem.</p> <p><strong>Remarks.</strong> </p> <ol> <li>Values $t\leq 2k$ may work only for convergents, not semiconvergents (the rational approximation in this case becomes so good that only convergents may satisfy it).</li> <li>There is a number of underwater stones here such as (i) not every set of values $a,b,c,\dots$ guarantee that we have a a continued fraction of the square root of an integer (in general, it's the square root of a rational); (ii) semiconvergents may not guarantee that the approximation is well enough for our purposes; etc. </li> </ol>
roots of higher derivatives of exponential
260,042
6
https://mathoverflow.net/questions/260042/roots-of-higher-derivatives-of-exponential
<p>Consider the Gaussian function $f(z)=e^{-z^2}$ which has no zeros on the complex domain. Let $D$ denote derivative w.r.t. the variable $z$.</p> <blockquote> <p><strong>Question.</strong> Is it true that $D^nf(z)=0$ has only real roots that are simple? If so, any slick proof?</p> </blockquote>
[ { "answer_id": 260047, "score": 19, "text": "<p>The (physicists') <a href=\"https://en.wikipedia.org/wiki/Hermite_polynomials\" rel=\"nofollow noreferrer\">Hermite polynomials</a> are</p>\n\n<p>$$ H_n(x) = (-1)^n e^{x^2} D^n e^{-x^2}$$</p>\n\n<p>And their roots are real. For that you don't need to know they are Hermite polynomials: just Rolle's theorem. See <a href=\"https://math.stackexchange.com/questions/104845/the-roots-of-hermite-polynomials-are-all-real\">this</a>.</p>\n" }, { "answer_id": 260051, "score": 10, "text": "<p>I've found a slightly different argument.</p>\n\n<p>If $f_m(z)=\\left(1-\\frac{z^2}m\\right)^m$ then $\\lim_{m\\rightarrow\\infty}f_m(z)=e^{-z^2}$ uniformly on every compact subset of $\\mathbb{C}$. Hence, the same holds for $\\lim_{m\\rightarrow\\infty}D^nf_m(z)=D^ne^{-z^2}$. On the other hand, $D^nf_m(z)$ has only real zeros. Therefore $D^ne^{-z^2}$ can not have non-real zeros, by <a href=\"https://en.wikipedia.org/wiki/Hurwitz%27s_theorem_(complex_analysis)\" rel=\"nofollow noreferrer\">Hurwitz's Theorem</a>.</p>\n" }, { "answer_id": 260079, "score": 9, "text": "<p>There are several complete characterizations of real entire functions whose all derivatives have all roots real: a) this is a closure of polynomials with the same property,\nand b) this class is represented by the formula\n$$f(z)=cz^me^{-az^2+bz}\\prod_{k}\\left(1-\\frac{z}{z_k}\\right)e^{z/z_k},$$\nwhere $a\\geq0$, $b,c$ are real, $m\\geq 0$ is an integer and $z_k$ real, with\n$$\\sum\\frac{1}{|z_k|^2}&lt;\\infty.$$\nThis is a parametric description: each such function is represented by this formula,\nand each function represented by this formula has the stated property.\nThis class of function has a standard name: Polya-Wiman class.</p>\n\n<p>These results are due to Wiman, Polya, Hellerstein and Williamson.\nFor recent generalizations and survey, see\n arXiv:math/0510502.</p>\n" }, { "answer_id": 386704, "score": 0, "text": "<p>There are several ways to skin this cat. Most are presented in the excellent survey &quot;<a href=\"https://www.dcce.ibilce.unesp.br/%7Edimitrov/ProfRusev/EJA_Paper_24_02_11/Dim_Rus_main.pdf\" rel=\"nofollow noreferrer\">Zeros of entire Fourier transforms</a>&quot; by\nDimitar K. Dimitrov and Peter K. Rusev with an account of the historical development. Jensen's contributions are particularly interesting to me with the emphasis on Appell polynomial sequences--the Hermite polynomials are the only set of orthogonal Appell polynomials.</p>\n" }, { "answer_id": 260047, "score": 19, "text": "<p>The (physicists') <a href=\"https://en.wikipedia.org/wiki/Hermite_polynomials\" rel=\"nofollow noreferrer\">Hermite polynomials</a> are</p>\n\n<p>$$ H_n(x) = (-1)^n e^{x^2} D^n e^{-x^2}$$</p>\n\n<p>And their roots are real. For that you don't need to know they are Hermite polynomials: just Rolle's theorem. See <a href=\"https://math.stackexchange.com/questions/104845/the-roots-of-hermite-polynomials-are-all-real\">this</a>.</p>\n" }, { "answer_id": 260051, "score": 10, "text": "<p>I've found a slightly different argument.</p>\n\n<p>If $f_m(z)=\\left(1-\\frac{z^2}m\\right)^m$ then $\\lim_{m\\rightarrow\\infty}f_m(z)=e^{-z^2}$ uniformly on every compact subset of $\\mathbb{C}$. Hence, the same holds for $\\lim_{m\\rightarrow\\infty}D^nf_m(z)=D^ne^{-z^2}$. On the other hand, $D^nf_m(z)$ has only real zeros. Therefore $D^ne^{-z^2}$ can not have non-real zeros, by <a href=\"https://en.wikipedia.org/wiki/Hurwitz%27s_theorem_(complex_analysis)\" rel=\"nofollow noreferrer\">Hurwitz's Theorem</a>.</p>\n" }, { "answer_id": 260079, "score": 9, "text": "<p>There are several complete characterizations of real entire functions whose all derivatives have all roots real: a) this is a closure of polynomials with the same property,\nand b) this class is represented by the formula\n$$f(z)=cz^me^{-az^2+bz}\\prod_{k}\\left(1-\\frac{z}{z_k}\\right)e^{z/z_k},$$\nwhere $a\\geq0$, $b,c$ are real, $m\\geq 0$ is an integer and $z_k$ real, with\n$$\\sum\\frac{1}{|z_k|^2}&lt;\\infty.$$\nThis is a parametric description: each such function is represented by this formula,\nand each function represented by this formula has the stated property.\nThis class of function has a standard name: Polya-Wiman class.</p>\n\n<p>These results are due to Wiman, Polya, Hellerstein and Williamson.\nFor recent generalizations and survey, see\n arXiv:math/0510502.</p>\n" }, { "answer_id": 386704, "score": 0, "text": "<p>There are several ways to skin this cat. Most are presented in the excellent survey &quot;<a href=\"https://www.dcce.ibilce.unesp.br/%7Edimitrov/ProfRusev/EJA_Paper_24_02_11/Dim_Rus_main.pdf\" rel=\"nofollow noreferrer\">Zeros of entire Fourier transforms</a>&quot; by\nDimitar K. Dimitrov and Peter K. Rusev with an account of the historical development. Jensen's contributions are particularly interesting to me with the emphasis on Appell polynomial sequences--the Hermite polynomials are the only set of orthogonal Appell polynomials.</p>\n" } ]
Question: roots of higher derivatives of exponential Answer 1: <p>The (physicists') <a href="https://en.wikipedia.org/wiki/Hermite_polynomials" rel="nofollow noreferrer">Hermite polynomials</a> are</p> <p>$$ H_n(x) = (-1)^n e^{x^2} D^n e^{-x^2}$$</p> <p>And their roots are real. For that you don't need to know they are Hermite polynomials: just Rolle's theorem. See <a href="https://math.stackexchange.com/questions/104845/the-roots-of-hermite-polynomials-are-all-real">this</a>.</p> Answer 2: <p>I've found a slightly different argument.</p> <p>If $f_m(z)=\left(1-\frac{z^2}m\right)^m$ then $\lim_{m\rightarrow\infty}f_m(z)=e^{-z^2}$ uniformly on every compact subset of $\mathbb{C}$. Hence, the same holds for $\lim_{m\rightarrow\infty}D^nf_m(z)=D^ne^{-z^2}$. On the other hand, $D^nf_m(z)$ has only real zeros. Therefore $D^ne^{-z^2}$ can not have non-real zeros, by <a href="https://en.wikipedia.org/wiki/Hurwitz%27s_theorem_(complex_analysis)" rel="nofollow noreferrer">Hurwitz's Theorem</a>.</p> Answer 3: <p>There are several complete characterizations of real entire functions whose all derivatives have all roots real: a) this is a closure of polynomials with the same property, and b) this class is represented by the formula $$f(z)=cz^me^{-az^2+bz}\prod_{k}\left(1-\frac{z}{z_k}\right)e^{z/z_k},$$ where $a\geq0$, $b,c$ are real, $m\geq 0$ is an integer and $z_k$ real, with $$\sum\frac{1}{|z_k|^2}&lt;\infty.$$ This is a parametric description: each such function is represented by this formula, and each function represented by this formula has the stated property. This class of function has a standard name: Polya-Wiman class.</p> <p>These results are due to Wiman, Polya, Hellerstein and Williamson. For recent generalizations and survey, see arXiv:math/0510502.</p> Answer 4: <p>There are several ways to skin this cat. Most are presented in the excellent survey &quot;<a href="https://www.dcce.ibilce.unesp.br/%7Edimitrov/ProfRusev/EJA_Paper_24_02_11/Dim_Rus_main.pdf" rel="nofollow noreferrer">Zeros of entire Fourier transforms</a>&quot; by Dimitar K. Dimitrov and Peter K. Rusev with an account of the historical development. Jensen's contributions are particularly interesting to me with the emphasis on Appell polynomial sequences--the Hermite polynomials are the only set of orthogonal Appell polynomials.</p> Answer 5: <p>The (physicists') <a href="https://en.wikipedia.org/wiki/Hermite_polynomials" rel="nofollow noreferrer">Hermite polynomials</a> are</p> <p>$$ H_n(x) = (-1)^n e^{x^2} D^n e^{-x^2}$$</p> <p>And their roots are real. For that you don't need to know they are Hermite polynomials: just Rolle's theorem. See <a href="https://math.stackexchange.com/questions/104845/the-roots-of-hermite-polynomials-are-all-real">this</a>.</p> Answer 6: <p>I've found a slightly different argument.</p> <p>If $f_m(z)=\left(1-\frac{z^2}m\right)^m$ then $\lim_{m\rightarrow\infty}f_m(z)=e^{-z^2}$ uniformly on every compact subset of $\mathbb{C}$. Hence, the same holds for $\lim_{m\rightarrow\infty}D^nf_m(z)=D^ne^{-z^2}$. On the other hand, $D^nf_m(z)$ has only real zeros. Therefore $D^ne^{-z^2}$ can not have non-real zeros, by <a href="https://en.wikipedia.org/wiki/Hurwitz%27s_theorem_(complex_analysis)" rel="nofollow noreferrer">Hurwitz's Theorem</a>.</p> Answer 7: <p>There are several complete characterizations of real entire functions whose all derivatives have all roots real: a) this is a closure of polynomials with the same property, and b) this class is represented by the formula $$f(z)=cz^me^{-az^2+bz}\prod_{k}\left(1-\frac{z}{z_k}\right)e^{z/z_k},$$ where $a\geq0$, $b,c$ are real, $m\geq 0$ is an integer and $z_k$ real, with $$\sum\frac{1}{|z_k|^2}&lt;\infty.$$ This is a parametric description: each such function is represented by this formula, and each function represented by this formula has the stated property. This class of function has a standard name: Polya-Wiman class.</p> <p>These results are due to Wiman, Polya, Hellerstein and Williamson. For recent generalizations and survey, see arXiv:math/0510502.</p> Answer 8: <p>There are several ways to skin this cat. Most are presented in the excellent survey &quot;<a href="https://www.dcce.ibilce.unesp.br/%7Edimitrov/ProfRusev/EJA_Paper_24_02_11/Dim_Rus_main.pdf" rel="nofollow noreferrer">Zeros of entire Fourier transforms</a>&quot; by Dimitar K. Dimitrov and Peter K. Rusev with an account of the historical development. Jensen's contributions are particularly interesting to me with the emphasis on Appell polynomial sequences--the Hermite polynomials are the only set of orthogonal Appell polynomials.</p>
&quot;Joyal type&quot; model structure for (n,1)-categories?
260,106
6
https://mathoverflow.net/questions/260106/joyal-type-model-structure-for-n-1-categories
<p>The Joyal model structure on the category of simplicial sets, has monomorphisms as cofibrations and quasi-categories as fibrant objects (these model $(\infty,1)$-categories). In HTT (section 2.3.4) Lurie defines the notion of an $n$-category (these model $(n,1)$-categories) and proves some theorems about them. For example, $1$-categories are just simplicial sets which are isomorphic to a nerve of a category (prop. 2.3.4.5). Moreover, every quasi-category $\mathcal{C}$ has an $n$-truncation $h_n\mathcal{C}$ which is an $n$-category with a map $\mathcal{C} \to h_n\mathcal{C}$ which is universal from $\mathcal{C}$ to an $n$-category (see prop. 2.3.4.12 for a precise statement).</p> <blockquote> <p><strong>Qusetion:</strong> Is there a model structure on the category of simplicial sets with monomorphisms as cofibrations, $n$-categories as fibrant objects and such that two quasi-categories are equivalent in this model structure if and only if their $n$-truncations are equivalent in the Joyal model structure?</p> </blockquote> <p>I ask this mainly out of pure curiosity and haven't put too much effort in thinking about it myself, in the hope that someone already knows the answer. Hence, perhaps there are simple reasons that this can't happen, in this case (and any other case), I'm open to considering some variations. </p> <p><strong>Remark</strong>: In fact, as I learned from <a href="https://mathoverflow.net/questions/29635/what-determines-a-model-structure">this</a> question, the cofibrations and fibrant objects determine the model strucute, so the question can be broken in two. First, whether there is a model structure with the specified cofibrations and fibrant objects, and if so, how can we interpret the weak equivalences. </p>
[ { "answer_id": 260110, "score": 7, "text": "<p>I don't know if this will work with the definition of $(n, 1)$-categories exactly as stated by Lurie, but it will work with a reasonable modification.</p>\n\n<p>Definition 2.3.4.1 from HTT basically says that a quasicategory is an $(n, 1)$-category if it has no non-trivial morphisms above dimension $n$ and it insists that this condition is satisfied <em>on the nose</em>. This is probably too much to ask if we aim for a nice model for the homotopy theory of $(n, 1)$-categories. (In particular, the definition is not invariant under categorical equivalences.) We would rather impose this condition <em>homotopically</em>.</p>\n\n<p>It is perfectly sensible to define an $(n, 1)$-category as a quasicategory whose mapping spaces are $(n-1)$-truncated. This can be concisely stated as follows. An $(n, 1)$-category is a quasicategory $\\mathcal{C}$ with the right lifting property with respect to inclusions $\\partial\\Delta[m] \\to \\Delta[m]$ for all $m \\ge n + 2$. I'm pretty sure that this is equivalent to $\\mathcal{C}$ being categorically equivalent to an $(n, 1)$-category in the sense of HTT.</p>\n\n<p>Now we can simply take the left Bousfield localization of the Joyal model structure with respect to the maps $\\partial\\Delta[m] \\to \\Delta[m]$ for $m \\ge n + 2$. This exists by the general theory of Bousfield localizations (see e.g. Hirschhorn's book) and has $(n, 1)$-categories as fibrant objects, monomorphisms as cofibrations and weak equivalences between fibrant objects are exactly the categorical equivalences.</p>\n\n<p>I'm not aware of any reference that constructs this model category, much less one that develops it any further.</p>\n" }, { "answer_id": 345759, "score": 2, "text": "<p>You may be interested in having a look at this: <a href=\"https://arxiv.org/abs/1810.11188\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1810.11188</a>. </p>\n" }, { "answer_id": 260110, "score": 7, "text": "<p>I don't know if this will work with the definition of $(n, 1)$-categories exactly as stated by Lurie, but it will work with a reasonable modification.</p>\n\n<p>Definition 2.3.4.1 from HTT basically says that a quasicategory is an $(n, 1)$-category if it has no non-trivial morphisms above dimension $n$ and it insists that this condition is satisfied <em>on the nose</em>. This is probably too much to ask if we aim for a nice model for the homotopy theory of $(n, 1)$-categories. (In particular, the definition is not invariant under categorical equivalences.) We would rather impose this condition <em>homotopically</em>.</p>\n\n<p>It is perfectly sensible to define an $(n, 1)$-category as a quasicategory whose mapping spaces are $(n-1)$-truncated. This can be concisely stated as follows. An $(n, 1)$-category is a quasicategory $\\mathcal{C}$ with the right lifting property with respect to inclusions $\\partial\\Delta[m] \\to \\Delta[m]$ for all $m \\ge n + 2$. I'm pretty sure that this is equivalent to $\\mathcal{C}$ being categorically equivalent to an $(n, 1)$-category in the sense of HTT.</p>\n\n<p>Now we can simply take the left Bousfield localization of the Joyal model structure with respect to the maps $\\partial\\Delta[m] \\to \\Delta[m]$ for $m \\ge n + 2$. This exists by the general theory of Bousfield localizations (see e.g. Hirschhorn's book) and has $(n, 1)$-categories as fibrant objects, monomorphisms as cofibrations and weak equivalences between fibrant objects are exactly the categorical equivalences.</p>\n\n<p>I'm not aware of any reference that constructs this model category, much less one that develops it any further.</p>\n" }, { "answer_id": 345759, "score": 2, "text": "<p>You may be interested in having a look at this: <a href=\"https://arxiv.org/abs/1810.11188\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1810.11188</a>. </p>\n" } ]
Question: &quot;Joyal type&quot; model structure for (n,1)-categories? Answer 1: <p>I don't know if this will work with the definition of $(n, 1)$-categories exactly as stated by Lurie, but it will work with a reasonable modification.</p> <p>Definition 2.3.4.1 from HTT basically says that a quasicategory is an $(n, 1)$-category if it has no non-trivial morphisms above dimension $n$ and it insists that this condition is satisfied <em>on the nose</em>. This is probably too much to ask if we aim for a nice model for the homotopy theory of $(n, 1)$-categories. (In particular, the definition is not invariant under categorical equivalences.) We would rather impose this condition <em>homotopically</em>.</p> <p>It is perfectly sensible to define an $(n, 1)$-category as a quasicategory whose mapping spaces are $(n-1)$-truncated. This can be concisely stated as follows. An $(n, 1)$-category is a quasicategory $\mathcal{C}$ with the right lifting property with respect to inclusions $\partial\Delta[m] \to \Delta[m]$ for all $m \ge n + 2$. I'm pretty sure that this is equivalent to $\mathcal{C}$ being categorically equivalent to an $(n, 1)$-category in the sense of HTT.</p> <p>Now we can simply take the left Bousfield localization of the Joyal model structure with respect to the maps $\partial\Delta[m] \to \Delta[m]$ for $m \ge n + 2$. This exists by the general theory of Bousfield localizations (see e.g. Hirschhorn's book) and has $(n, 1)$-categories as fibrant objects, monomorphisms as cofibrations and weak equivalences between fibrant objects are exactly the categorical equivalences.</p> <p>I'm not aware of any reference that constructs this model category, much less one that develops it any further.</p> Answer 2: <p>You may be interested in having a look at this: <a href="https://arxiv.org/abs/1810.11188" rel="nofollow noreferrer">https://arxiv.org/abs/1810.11188</a>. </p> Answer 3: <p>I don't know if this will work with the definition of $(n, 1)$-categories exactly as stated by Lurie, but it will work with a reasonable modification.</p> <p>Definition 2.3.4.1 from HTT basically says that a quasicategory is an $(n, 1)$-category if it has no non-trivial morphisms above dimension $n$ and it insists that this condition is satisfied <em>on the nose</em>. This is probably too much to ask if we aim for a nice model for the homotopy theory of $(n, 1)$-categories. (In particular, the definition is not invariant under categorical equivalences.) We would rather impose this condition <em>homotopically</em>.</p> <p>It is perfectly sensible to define an $(n, 1)$-category as a quasicategory whose mapping spaces are $(n-1)$-truncated. This can be concisely stated as follows. An $(n, 1)$-category is a quasicategory $\mathcal{C}$ with the right lifting property with respect to inclusions $\partial\Delta[m] \to \Delta[m]$ for all $m \ge n + 2$. I'm pretty sure that this is equivalent to $\mathcal{C}$ being categorically equivalent to an $(n, 1)$-category in the sense of HTT.</p> <p>Now we can simply take the left Bousfield localization of the Joyal model structure with respect to the maps $\partial\Delta[m] \to \Delta[m]$ for $m \ge n + 2$. This exists by the general theory of Bousfield localizations (see e.g. Hirschhorn's book) and has $(n, 1)$-categories as fibrant objects, monomorphisms as cofibrations and weak equivalences between fibrant objects are exactly the categorical equivalences.</p> <p>I'm not aware of any reference that constructs this model category, much less one that develops it any further.</p> Answer 4: <p>You may be interested in having a look at this: <a href="https://arxiv.org/abs/1810.11188" rel="nofollow noreferrer">https://arxiv.org/abs/1810.11188</a>. </p>
Inner product of columns of a matrix
260,411
6
https://mathoverflow.net/questions/260411/inner-product-of-columns-of-a-matrix
<p>What information about a (square) matrix we earn if the inner product of its columns are known?</p>
[ { "answer_id": 260412, "score": 16, "text": "<p>You have determined $A^*A$, or, alternatively, you know $A$ up to pre-multiplication by a unitary matrix $U$. So you know the $R$ factor of its QR factorization, and the factors $\\Sigma$ and $V$ of its SVD. In particular, among other things, you know singular values and right singular vectors, the sign of its determinant, its rank, its kernel, its Euclidean and Frobenius norm.</p>\n" }, { "answer_id": 260429, "score": 8, "text": "<p>A more geometric supplement to Federico's answer: if you think of columns as vectors, then knowing the inner products determines the set of vectors up to rigid motion, so any geometric information is determined by the inner products (including the volume of the simplex they span [determinant], singular values [the semi-axes of the ellipsoid which is the image of the unit ball by your matrix], the areas of the faces of the simplex determined by them, etc, etc.</p>\n" }, { "answer_id": 260412, "score": 16, "text": "<p>You have determined $A^*A$, or, alternatively, you know $A$ up to pre-multiplication by a unitary matrix $U$. So you know the $R$ factor of its QR factorization, and the factors $\\Sigma$ and $V$ of its SVD. In particular, among other things, you know singular values and right singular vectors, the sign of its determinant, its rank, its kernel, its Euclidean and Frobenius norm.</p>\n" }, { "answer_id": 260429, "score": 8, "text": "<p>A more geometric supplement to Federico's answer: if you think of columns as vectors, then knowing the inner products determines the set of vectors up to rigid motion, so any geometric information is determined by the inner products (including the volume of the simplex they span [determinant], singular values [the semi-axes of the ellipsoid which is the image of the unit ball by your matrix], the areas of the faces of the simplex determined by them, etc, etc.</p>\n" } ]
Question: Inner product of columns of a matrix Answer 1: <p>You have determined $A^*A$, or, alternatively, you know $A$ up to pre-multiplication by a unitary matrix $U$. So you know the $R$ factor of its QR factorization, and the factors $\Sigma$ and $V$ of its SVD. In particular, among other things, you know singular values and right singular vectors, the sign of its determinant, its rank, its kernel, its Euclidean and Frobenius norm.</p> Answer 2: <p>A more geometric supplement to Federico's answer: if you think of columns as vectors, then knowing the inner products determines the set of vectors up to rigid motion, so any geometric information is determined by the inner products (including the volume of the simplex they span [determinant], singular values [the semi-axes of the ellipsoid which is the image of the unit ball by your matrix], the areas of the faces of the simplex determined by them, etc, etc.</p> Answer 3: <p>You have determined $A^*A$, or, alternatively, you know $A$ up to pre-multiplication by a unitary matrix $U$. So you know the $R$ factor of its QR factorization, and the factors $\Sigma$ and $V$ of its SVD. In particular, among other things, you know singular values and right singular vectors, the sign of its determinant, its rank, its kernel, its Euclidean and Frobenius norm.</p> Answer 4: <p>A more geometric supplement to Federico's answer: if you think of columns as vectors, then knowing the inner products determines the set of vectors up to rigid motion, so any geometric information is determined by the inner products (including the volume of the simplex they span [determinant], singular values [the semi-axes of the ellipsoid which is the image of the unit ball by your matrix], the areas of the faces of the simplex determined by them, etc, etc.</p>
A class of tilings with amazing visual qualities
260,567
6
https://mathoverflow.net/questions/260567/a-class-of-tilings-with-amazing-visual-qualities
<p>For more examples please see my related question on MSE: <a href="https://math.stackexchange.com/q/2045725/266335">Interesting tiling with a lot of symmetrical shapes</a> </p> <p><a href="https://i.stack.imgur.com/bSSJ2.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/bSSJ2.png" alt="enter image description here"></a> </p> <p>This is achieved by rotation of square grid over itself by atan(3/4). Resulting grid is a base for a whole class of beautiful tilings, see examples in the linked question. Playing with this grid I can make up many tilesets and most interesting is, that they can act in 8 directions (4 axes). Namely I can represent contiguous equal-width lines in all 8 directions and respectively all shapes formed by these lines. Resulting shapes have smooth transitions, no gaps and are 'scalable'.<br> On the other hand such tilesets are very simple and with low amount of unique tiles. So in other words, it is a whole nugget of interesting applications, e.g. in vision and typography.</p> <p>Still I was not able to find any concrete source related to this tiling class. </p> <p>So can you recommend any good sources dedicated to this phenomenon? </p> <hr> <p><strong>Update</strong>:</p> <p>It has some connection to <a href="https://en.wikipedia.org/wiki/Pythagorean_tiling" rel="nofollow noreferrer">Pythagorean tiling</a>, namely a Pythagorean tiling with 1:2 square ratio can be used to construct this tiling. </p> <p>Draw lines through centers of big squares in such manner: <a href="https://i.stack.imgur.com/Av4vW.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/Av4vW.png" alt="enter image description here"></a></p> <p>After drawing lines all over the plane it results in the above described tiling. Here it is painted in two colors:</p> <p><a href="https://i.stack.imgur.com/IN4lS.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/IN4lS.png" alt="enter image description here"></a></p>
[ { "answer_id": 261463, "score": 5, "text": "<p>These multigrids can be used to construct Penrose tilings. Here we are overlaying five sets of parallel lines. And I will stop here because you can read the article, using the grids alone try to fill out the tiling. </p>\n\n<p><a href=\"http://www.ams.org/samplings/feature-column/fcarc-ribbons\" rel=\"nofollow noreferrer\">http://www.ams.org/samplings/feature-column/fcarc-ribbons</a></p>\n\n<p><a href=\"https://i.stack.imgur.com/eLsJl.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/eLsJl.jpg\" alt=\"enter image description here\"></a></p>\n\n<p>now MathOverflow is a research site but I will advocate this is a research level topic</p>\n\n<ul>\n<li><p>Penrose tilings are mentioned as an example of a \"non-commutative\" space in the book of Alain Connes.</p></li>\n<li><p>The spectra of these grids also have some pecular and unexplored properties. Here is an example from <a href=\"https://arxiv.org/abs/1501.01198\" rel=\"nofollow noreferrer\">2015</a> </p></li>\n<li><p>This paper combines properties of visible points and Penrose tilings (from <a href=\"https://arxiv.org/abs/1404.1564\" rel=\"nofollow noreferrer\">2014</a>)</p></li>\n</ul>\n\n<p><a href=\"https://i.stack.imgur.com/SAZWT.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/SAZWT.png\" alt=\"enter image description here\"></a></p>\n\n<h3>Amman-Beenker Tilings (Tetragrid)</h3>\n\n<p><a href=\"https://i.stack.imgur.com/es5c7.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/es5c7.png\" alt=\"enter image description here\"></a></p>\n\n<p>Beenker's original paper <a href=\"http://repository.tue.nl/253166\" rel=\"nofollow noreferrer\">Algebraic theory of non-periodic tilings of the plane by two simple building blocks : a square and a rhombus</a> is online. And he shows you how to use the tetragrid to build the squares and rhombuses one by one.</p>\n\n<p><a href=\"https://i.stack.imgur.com/V7JiJ.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/V7JiJ.png\" alt=\"enter image description here\"></a></p>\n\n<h3>Muslim Architecture and Design</h3>\n\n<p>Tiles with 8- and 10-fold symmetry appear in Muslim (?) architecture and design as well as Arabic-influenced Spanish architecture</p>\n\n<p><img src=\"https://upload.wikimedia.org/wikipedia/commons/8/83/Green_mosque_archway.JPG\" alt=\"\"></p>\n" }, { "answer_id": 261845, "score": 0, "text": "<p>This appears to be related to the topic of tilings with multiplicity, i.e. where almost every point is covered by fixed number $k$ of tiles, see <a href=\"https://arxiv.org/abs/1103.3163\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1103.3163</a>, \n<a href=\"https://arxiv.org/abs/1405.6928\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1405.6928</a> and references therein.</p>\n" }, { "answer_id": 261463, "score": 5, "text": "<p>These multigrids can be used to construct Penrose tilings. Here we are overlaying five sets of parallel lines. And I will stop here because you can read the article, using the grids alone try to fill out the tiling. </p>\n\n<p><a href=\"http://www.ams.org/samplings/feature-column/fcarc-ribbons\" rel=\"nofollow noreferrer\">http://www.ams.org/samplings/feature-column/fcarc-ribbons</a></p>\n\n<p><a href=\"https://i.stack.imgur.com/eLsJl.jpg\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/eLsJl.jpg\" alt=\"enter image description here\"></a></p>\n\n<p>now MathOverflow is a research site but I will advocate this is a research level topic</p>\n\n<ul>\n<li><p>Penrose tilings are mentioned as an example of a \"non-commutative\" space in the book of Alain Connes.</p></li>\n<li><p>The spectra of these grids also have some pecular and unexplored properties. Here is an example from <a href=\"https://arxiv.org/abs/1501.01198\" rel=\"nofollow noreferrer\">2015</a> </p></li>\n<li><p>This paper combines properties of visible points and Penrose tilings (from <a href=\"https://arxiv.org/abs/1404.1564\" rel=\"nofollow noreferrer\">2014</a>)</p></li>\n</ul>\n\n<p><a href=\"https://i.stack.imgur.com/SAZWT.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/SAZWT.png\" alt=\"enter image description here\"></a></p>\n\n<h3>Amman-Beenker Tilings (Tetragrid)</h3>\n\n<p><a href=\"https://i.stack.imgur.com/es5c7.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/es5c7.png\" alt=\"enter image description here\"></a></p>\n\n<p>Beenker's original paper <a href=\"http://repository.tue.nl/253166\" rel=\"nofollow noreferrer\">Algebraic theory of non-periodic tilings of the plane by two simple building blocks : a square and a rhombus</a> is online. And he shows you how to use the tetragrid to build the squares and rhombuses one by one.</p>\n\n<p><a href=\"https://i.stack.imgur.com/V7JiJ.png\" rel=\"nofollow noreferrer\"><img src=\"https://i.stack.imgur.com/V7JiJ.png\" alt=\"enter image description here\"></a></p>\n\n<h3>Muslim Architecture and Design</h3>\n\n<p>Tiles with 8- and 10-fold symmetry appear in Muslim (?) architecture and design as well as Arabic-influenced Spanish architecture</p>\n\n<p><img src=\"https://upload.wikimedia.org/wikipedia/commons/8/83/Green_mosque_archway.JPG\" alt=\"\"></p>\n" }, { "answer_id": 261845, "score": 0, "text": "<p>This appears to be related to the topic of tilings with multiplicity, i.e. where almost every point is covered by fixed number $k$ of tiles, see <a href=\"https://arxiv.org/abs/1103.3163\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1103.3163</a>, \n<a href=\"https://arxiv.org/abs/1405.6928\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1405.6928</a> and references therein.</p>\n" } ]
Question: A class of tilings with amazing visual qualities Answer 1: <p>These multigrids can be used to construct Penrose tilings. Here we are overlaying five sets of parallel lines. And I will stop here because you can read the article, using the grids alone try to fill out the tiling. </p> <p><a href="http://www.ams.org/samplings/feature-column/fcarc-ribbons" rel="nofollow noreferrer">http://www.ams.org/samplings/feature-column/fcarc-ribbons</a></p> <p><a href="https://i.stack.imgur.com/eLsJl.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/eLsJl.jpg" alt="enter image description here"></a></p> <p>now MathOverflow is a research site but I will advocate this is a research level topic</p> <ul> <li><p>Penrose tilings are mentioned as an example of a "non-commutative" space in the book of Alain Connes.</p></li> <li><p>The spectra of these grids also have some pecular and unexplored properties. Here is an example from <a href="https://arxiv.org/abs/1501.01198" rel="nofollow noreferrer">2015</a> </p></li> <li><p>This paper combines properties of visible points and Penrose tilings (from <a href="https://arxiv.org/abs/1404.1564" rel="nofollow noreferrer">2014</a>)</p></li> </ul> <p><a href="https://i.stack.imgur.com/SAZWT.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/SAZWT.png" alt="enter image description here"></a></p> <h3>Amman-Beenker Tilings (Tetragrid)</h3> <p><a href="https://i.stack.imgur.com/es5c7.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/es5c7.png" alt="enter image description here"></a></p> <p>Beenker's original paper <a href="http://repository.tue.nl/253166" rel="nofollow noreferrer">Algebraic theory of non-periodic tilings of the plane by two simple building blocks : a square and a rhombus</a> is online. And he shows you how to use the tetragrid to build the squares and rhombuses one by one.</p> <p><a href="https://i.stack.imgur.com/V7JiJ.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/V7JiJ.png" alt="enter image description here"></a></p> <h3>Muslim Architecture and Design</h3> <p>Tiles with 8- and 10-fold symmetry appear in Muslim (?) architecture and design as well as Arabic-influenced Spanish architecture</p> <p><img src="https://upload.wikimedia.org/wikipedia/commons/8/83/Green_mosque_archway.JPG" alt=""></p> Answer 2: <p>This appears to be related to the topic of tilings with multiplicity, i.e. where almost every point is covered by fixed number $k$ of tiles, see <a href="https://arxiv.org/abs/1103.3163" rel="nofollow noreferrer">https://arxiv.org/abs/1103.3163</a>, <a href="https://arxiv.org/abs/1405.6928" rel="nofollow noreferrer">https://arxiv.org/abs/1405.6928</a> and references therein.</p> Answer 3: <p>These multigrids can be used to construct Penrose tilings. Here we are overlaying five sets of parallel lines. And I will stop here because you can read the article, using the grids alone try to fill out the tiling. </p> <p><a href="http://www.ams.org/samplings/feature-column/fcarc-ribbons" rel="nofollow noreferrer">http://www.ams.org/samplings/feature-column/fcarc-ribbons</a></p> <p><a href="https://i.stack.imgur.com/eLsJl.jpg" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/eLsJl.jpg" alt="enter image description here"></a></p> <p>now MathOverflow is a research site but I will advocate this is a research level topic</p> <ul> <li><p>Penrose tilings are mentioned as an example of a "non-commutative" space in the book of Alain Connes.</p></li> <li><p>The spectra of these grids also have some pecular and unexplored properties. Here is an example from <a href="https://arxiv.org/abs/1501.01198" rel="nofollow noreferrer">2015</a> </p></li> <li><p>This paper combines properties of visible points and Penrose tilings (from <a href="https://arxiv.org/abs/1404.1564" rel="nofollow noreferrer">2014</a>)</p></li> </ul> <p><a href="https://i.stack.imgur.com/SAZWT.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/SAZWT.png" alt="enter image description here"></a></p> <h3>Amman-Beenker Tilings (Tetragrid)</h3> <p><a href="https://i.stack.imgur.com/es5c7.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/es5c7.png" alt="enter image description here"></a></p> <p>Beenker's original paper <a href="http://repository.tue.nl/253166" rel="nofollow noreferrer">Algebraic theory of non-periodic tilings of the plane by two simple building blocks : a square and a rhombus</a> is online. And he shows you how to use the tetragrid to build the squares and rhombuses one by one.</p> <p><a href="https://i.stack.imgur.com/V7JiJ.png" rel="nofollow noreferrer"><img src="https://i.stack.imgur.com/V7JiJ.png" alt="enter image description here"></a></p> <h3>Muslim Architecture and Design</h3> <p>Tiles with 8- and 10-fold symmetry appear in Muslim (?) architecture and design as well as Arabic-influenced Spanish architecture</p> <p><img src="https://upload.wikimedia.org/wikipedia/commons/8/83/Green_mosque_archway.JPG" alt=""></p> Answer 4: <p>This appears to be related to the topic of tilings with multiplicity, i.e. where almost every point is covered by fixed number $k$ of tiles, see <a href="https://arxiv.org/abs/1103.3163" rel="nofollow noreferrer">https://arxiv.org/abs/1103.3163</a>, <a href="https://arxiv.org/abs/1405.6928" rel="nofollow noreferrer">https://arxiv.org/abs/1405.6928</a> and references therein.</p>
Riemann Theta Function On Hyperbolic Riemann Surfaces
260,651
6
https://mathoverflow.net/questions/260651/riemann-theta-function-on-hyperbolic-riemann-surfaces
<p>The Riemann theta function for a genus $g$ closed Riemann surface with period matrix $\tau=[\tau_{ij}]$ is defined by</p> <p>$$\theta(\{z_1,\cdots,z_g\}|\tau)=\Sigma_{n\in\mathbb{Z}^g}e^{\pi i(n\cdot\tau\cdot n + 2n\cdot z)}$$</p> <p>The zeroth of this function is given by theta divisor, a divisor class $\Theta=W_{g-1}+\mathcal{K}$ in which $W_{g-1}$ is a degree $g-1$ divisor and $\mathcal{K}$ is the Riemann constant vector. I have two questions:</p> <ul> <li>How can Riemann theta function be defined for a hyperbolic Riemann surface?</li> <li>How can the theta divisor i.e. the locus of the zeroes of Riemann theta function be characterized in hyperbolic geometry? More specifically, Is there a way to characterize the theta divisor in terms of Fenchel-Nielsen coordinates on the Teichmuller space of hyperbolic Riemann surfaces?</li> </ul>
[ { "answer_id": 271048, "score": 6, "text": "<p>The answer to the first question is easy: The hyperbolic metric defines a complex structure, and the Theta function only depends on the complex structure, so take that one. </p>\n\n<p>The second question is <strong>much</strong> harder as it relates the algebraic/complex world (Jacobian, Theta function) with the trancendental/real-analytic world (the hyperbolic metric). </p>\n\n<p>There are asymptotic results, however, relating pinching of curves to the periods of abelian differentials (starting the older work of Rauch, Masur and Royden). Check the keywords Rauch variation formulas, Satake-Baily-Borel compactification of moduli space, asymptotics of the L_2-Bergman metric, etc.</p>\n" }, { "answer_id": 271073, "score": 2, "text": "<p>To amplify on ElucisusFTW's answer, the relationship between the hyperbolic metric and the complex structure is the so-called \"accessory parameters\" problem, and very little is known about it, except in the very special setting of hyperbolic punctured spheres (and the related punctured tori), where there is a large number of papers by Zograf and Takhtajan (warning, the latter name is spelled in many different ways). This is closely related also to string theory (which I know next to nothing about), and the determinants of Laplacians, which CAN be computed in terms of hyperbolic invariants. See, for example, Pollicott-Rocha and Strohmaier-Uski (note that the latter paper says that the former paper is wrong...)</p>\n\n<p><em>Pollicott, Mark; Rocha, Andr\\'e C.</em>, <a href=\"http://dx.doi.org/10.1007/s002220050190\" rel=\"nofollow noreferrer\"><strong>A remarkable formula for the determinant of the Laplacian</strong></a>, Invent. Math. 130, No.2, 399-414 (1997). <a href=\"https://zbmath.org/?q=an:0896.58067\" rel=\"nofollow noreferrer\">ZBL0896.58067</a>.</p>\n\n<p><em>Strohmaier, Alexander; Uski, Ville</em>, <a href=\"http://dx.doi.org/10.1007/s00220-012-1557-1\" rel=\"nofollow noreferrer\"><strong>An algorithm for the computation of eigenvalues, spectral zeta functions and zeta-determinants on hyperbolic surfaces</strong></a>, Commun. Math. Phys. 317, No. 3, 827-869 (2013). <a href=\"https://zbmath.org/?q=an:1261.65113\" rel=\"nofollow noreferrer\">ZBL1261.65113</a>.</p>\n" }, { "answer_id": 271048, "score": 6, "text": "<p>The answer to the first question is easy: The hyperbolic metric defines a complex structure, and the Theta function only depends on the complex structure, so take that one. </p>\n\n<p>The second question is <strong>much</strong> harder as it relates the algebraic/complex world (Jacobian, Theta function) with the trancendental/real-analytic world (the hyperbolic metric). </p>\n\n<p>There are asymptotic results, however, relating pinching of curves to the periods of abelian differentials (starting the older work of Rauch, Masur and Royden). Check the keywords Rauch variation formulas, Satake-Baily-Borel compactification of moduli space, asymptotics of the L_2-Bergman metric, etc.</p>\n" }, { "answer_id": 271073, "score": 2, "text": "<p>To amplify on ElucisusFTW's answer, the relationship between the hyperbolic metric and the complex structure is the so-called \"accessory parameters\" problem, and very little is known about it, except in the very special setting of hyperbolic punctured spheres (and the related punctured tori), where there is a large number of papers by Zograf and Takhtajan (warning, the latter name is spelled in many different ways). This is closely related also to string theory (which I know next to nothing about), and the determinants of Laplacians, which CAN be computed in terms of hyperbolic invariants. See, for example, Pollicott-Rocha and Strohmaier-Uski (note that the latter paper says that the former paper is wrong...)</p>\n\n<p><em>Pollicott, Mark; Rocha, Andr\\'e C.</em>, <a href=\"http://dx.doi.org/10.1007/s002220050190\" rel=\"nofollow noreferrer\"><strong>A remarkable formula for the determinant of the Laplacian</strong></a>, Invent. Math. 130, No.2, 399-414 (1997). <a href=\"https://zbmath.org/?q=an:0896.58067\" rel=\"nofollow noreferrer\">ZBL0896.58067</a>.</p>\n\n<p><em>Strohmaier, Alexander; Uski, Ville</em>, <a href=\"http://dx.doi.org/10.1007/s00220-012-1557-1\" rel=\"nofollow noreferrer\"><strong>An algorithm for the computation of eigenvalues, spectral zeta functions and zeta-determinants on hyperbolic surfaces</strong></a>, Commun. Math. Phys. 317, No. 3, 827-869 (2013). <a href=\"https://zbmath.org/?q=an:1261.65113\" rel=\"nofollow noreferrer\">ZBL1261.65113</a>.</p>\n" } ]
Question: Riemann Theta Function On Hyperbolic Riemann Surfaces Answer 1: <p>The answer to the first question is easy: The hyperbolic metric defines a complex structure, and the Theta function only depends on the complex structure, so take that one. </p> <p>The second question is <strong>much</strong> harder as it relates the algebraic/complex world (Jacobian, Theta function) with the trancendental/real-analytic world (the hyperbolic metric). </p> <p>There are asymptotic results, however, relating pinching of curves to the periods of abelian differentials (starting the older work of Rauch, Masur and Royden). Check the keywords Rauch variation formulas, Satake-Baily-Borel compactification of moduli space, asymptotics of the L_2-Bergman metric, etc.</p> Answer 2: <p>To amplify on ElucisusFTW's answer, the relationship between the hyperbolic metric and the complex structure is the so-called "accessory parameters" problem, and very little is known about it, except in the very special setting of hyperbolic punctured spheres (and the related punctured tori), where there is a large number of papers by Zograf and Takhtajan (warning, the latter name is spelled in many different ways). This is closely related also to string theory (which I know next to nothing about), and the determinants of Laplacians, which CAN be computed in terms of hyperbolic invariants. See, for example, Pollicott-Rocha and Strohmaier-Uski (note that the latter paper says that the former paper is wrong...)</p> <p><em>Pollicott, Mark; Rocha, Andr\'e C.</em>, <a href="http://dx.doi.org/10.1007/s002220050190" rel="nofollow noreferrer"><strong>A remarkable formula for the determinant of the Laplacian</strong></a>, Invent. Math. 130, No.2, 399-414 (1997). <a href="https://zbmath.org/?q=an:0896.58067" rel="nofollow noreferrer">ZBL0896.58067</a>.</p> <p><em>Strohmaier, Alexander; Uski, Ville</em>, <a href="http://dx.doi.org/10.1007/s00220-012-1557-1" rel="nofollow noreferrer"><strong>An algorithm for the computation of eigenvalues, spectral zeta functions and zeta-determinants on hyperbolic surfaces</strong></a>, Commun. Math. Phys. 317, No. 3, 827-869 (2013). <a href="https://zbmath.org/?q=an:1261.65113" rel="nofollow noreferrer">ZBL1261.65113</a>.</p> Answer 3: <p>The answer to the first question is easy: The hyperbolic metric defines a complex structure, and the Theta function only depends on the complex structure, so take that one. </p> <p>The second question is <strong>much</strong> harder as it relates the algebraic/complex world (Jacobian, Theta function) with the trancendental/real-analytic world (the hyperbolic metric). </p> <p>There are asymptotic results, however, relating pinching of curves to the periods of abelian differentials (starting the older work of Rauch, Masur and Royden). Check the keywords Rauch variation formulas, Satake-Baily-Borel compactification of moduli space, asymptotics of the L_2-Bergman metric, etc.</p> Answer 4: <p>To amplify on ElucisusFTW's answer, the relationship between the hyperbolic metric and the complex structure is the so-called "accessory parameters" problem, and very little is known about it, except in the very special setting of hyperbolic punctured spheres (and the related punctured tori), where there is a large number of papers by Zograf and Takhtajan (warning, the latter name is spelled in many different ways). This is closely related also to string theory (which I know next to nothing about), and the determinants of Laplacians, which CAN be computed in terms of hyperbolic invariants. See, for example, Pollicott-Rocha and Strohmaier-Uski (note that the latter paper says that the former paper is wrong...)</p> <p><em>Pollicott, Mark; Rocha, Andr\'e C.</em>, <a href="http://dx.doi.org/10.1007/s002220050190" rel="nofollow noreferrer"><strong>A remarkable formula for the determinant of the Laplacian</strong></a>, Invent. Math. 130, No.2, 399-414 (1997). <a href="https://zbmath.org/?q=an:0896.58067" rel="nofollow noreferrer">ZBL0896.58067</a>.</p> <p><em>Strohmaier, Alexander; Uski, Ville</em>, <a href="http://dx.doi.org/10.1007/s00220-012-1557-1" rel="nofollow noreferrer"><strong>An algorithm for the computation of eigenvalues, spectral zeta functions and zeta-determinants on hyperbolic surfaces</strong></a>, Commun. Math. Phys. 317, No. 3, 827-869 (2013). <a href="https://zbmath.org/?q=an:1261.65113" rel="nofollow noreferrer">ZBL1261.65113</a>.</p>
minimum-maximum entries matrix
258,752
5
https://mathoverflow.net/questions/258752/minimum-maximum-entries-matrix
<p>Let <span class="math-container">$M(n)$</span> be an <span class="math-container">$n\times n$</span> matrix in the variables <span class="math-container">$x_1,\dots,x_n$</span> with entries <span class="math-container">$$M_{i,j}(n)=\frac{x_{\max(i,j)}}{x_{\min(i,j)}}, \qquad 1\leq i,j\leq n.$$</span> I'm interested in the following:</p> <blockquote> <p><strong>Questions.</strong></p> <p>(1) Is there a neat or &quot;closed form&quot; evaluation for the determinant <span class="math-container">$\det M(n)$</span>?</p> <p>(2) Is there an explicit formula for the inverse of <span class="math-container">$M(n)$</span>?</p> </blockquote> <p>Thank you.</p>
[ { "answer_id": 258755, "score": 10, "text": "<p>Let us write $$a_r=\\frac{x_{r+1}}{x_r}$$ for $r=1\\cdots n$. </p>\n\n<p>We can then write the matrix $M(n)$ in the form \n$$\\begin{pmatrix} 1 &amp; a_1 &amp; a_1a_2&amp; \\cdots &amp; a_1a_2\\cdots a_{n-1} \\\\ a_1 &amp; 1 &amp; a_2&amp; \\cdots &amp; a_2\\cdots a_{n-1}\\\\ \\vdots &amp; \\vdots &amp; \\vdots &amp;\\ddots &amp; \\vdots\\\\ a_1a_2\\cdots a_{n-1}&amp; a_2\\cdots a_{n-1} &amp;\\cdots &amp; a_{n-1} &amp; 1\\end{pmatrix} $$ \nWe do now Gauss elimination, and reduce $a_{n-1}$ times the $(n-1)$-th row from the $n$-th row. We then get in the last row $0, 0, \\ldots (1-a_{n-1}^2)$.\nBut this means that $det(M(n))=det(M(n-1))(1-a_{n-1}^2)$ and by induction\n$$det(M(n)) = \\prod_{r=1}^{n-1} (1-a_r^2)=\\prod_{r=1}^{n-1}(1-\\frac{x_{r+1}^2}{x_r^2}).$$\nBy using inductively the Gauss elimination mentioned above, one can also get the inverse of $M(n)$.</p>\n" }, { "answer_id": 258769, "score": 3, "text": "<p>I wish to advertise the <a href=\"https://en.wikipedia.org/wiki/Dodgson_condensation\" rel=\"nofollow noreferrer\">Method of Condensation</a> in proving determinantal evaluation. The <strong>key</strong> is guess the answer, which is thanks to Ehud Meir. Then, generalize it a bit. Let $x_1, x_2,\\dots$ be an infinite set of variables and modify the original matrix (by shifting variables) to $M^{a,b}(n)$ so that\n$$M_{i,j}^{a,b}(n):=\\frac{x_{\\max(i+a,j+b)}}{x_{\\min(i+a,j+b)}}.$$\n<em>Convention:</em> $M^{0,0}(n)=M(n)$. </p>\n\n<p><strong>Claim.</strong> If $a\\neq b$ then $\\det M^{a,b}(n)=0$, and if $a=b$ then\n$$\\det M^{a,a}(n)=\\prod_{r=2}^n\\frac{x_{k-1+a}^2-x_{k+a}^2}{x_{k-1+a}^2}.\\tag1$$\n<strong>Proof.</strong> The case $a\\neq b$ is easy - simply factor out a variable from $n^{th}$-column/row and another variable from the $(n-1)^{th}$-column/row. These new columns/rows are identical.</p>\n\n<p>Inductive proofs neatly work with this Dodgson's recursive relation\n$$\\det Z^{0,0}(n)=\\frac{\\det Z^{1,1}(n-1)\\det Z^{1,1}(n-1)-\\det Z^{0,1}(n-1)\\det Z^{1,0}(n-1)}{\\det Z^{1,1}(n-2)}$$\nsatisfied by any matrix (so long as denominators do not vanish). Thus, it holds for $\\det M^{a,b}(n)$.\nSo, it remains to prove that the (explicit) formula on the RHS of (1) <em>does</em> satisfy the same equation. However, this is quite a routine simplification (preferably with symbolic sofwares). The proof follows. $\\square$</p>\n" }, { "answer_id": 258755, "score": 10, "text": "<p>Let us write $$a_r=\\frac{x_{r+1}}{x_r}$$ for $r=1\\cdots n$. </p>\n\n<p>We can then write the matrix $M(n)$ in the form \n$$\\begin{pmatrix} 1 &amp; a_1 &amp; a_1a_2&amp; \\cdots &amp; a_1a_2\\cdots a_{n-1} \\\\ a_1 &amp; 1 &amp; a_2&amp; \\cdots &amp; a_2\\cdots a_{n-1}\\\\ \\vdots &amp; \\vdots &amp; \\vdots &amp;\\ddots &amp; \\vdots\\\\ a_1a_2\\cdots a_{n-1}&amp; a_2\\cdots a_{n-1} &amp;\\cdots &amp; a_{n-1} &amp; 1\\end{pmatrix} $$ \nWe do now Gauss elimination, and reduce $a_{n-1}$ times the $(n-1)$-th row from the $n$-th row. We then get in the last row $0, 0, \\ldots (1-a_{n-1}^2)$.\nBut this means that $det(M(n))=det(M(n-1))(1-a_{n-1}^2)$ and by induction\n$$det(M(n)) = \\prod_{r=1}^{n-1} (1-a_r^2)=\\prod_{r=1}^{n-1}(1-\\frac{x_{r+1}^2}{x_r^2}).$$\nBy using inductively the Gauss elimination mentioned above, one can also get the inverse of $M(n)$.</p>\n" }, { "answer_id": 258769, "score": 3, "text": "<p>I wish to advertise the <a href=\"https://en.wikipedia.org/wiki/Dodgson_condensation\" rel=\"nofollow noreferrer\">Method of Condensation</a> in proving determinantal evaluation. The <strong>key</strong> is guess the answer, which is thanks to Ehud Meir. Then, generalize it a bit. Let $x_1, x_2,\\dots$ be an infinite set of variables and modify the original matrix (by shifting variables) to $M^{a,b}(n)$ so that\n$$M_{i,j}^{a,b}(n):=\\frac{x_{\\max(i+a,j+b)}}{x_{\\min(i+a,j+b)}}.$$\n<em>Convention:</em> $M^{0,0}(n)=M(n)$. </p>\n\n<p><strong>Claim.</strong> If $a\\neq b$ then $\\det M^{a,b}(n)=0$, and if $a=b$ then\n$$\\det M^{a,a}(n)=\\prod_{r=2}^n\\frac{x_{k-1+a}^2-x_{k+a}^2}{x_{k-1+a}^2}.\\tag1$$\n<strong>Proof.</strong> The case $a\\neq b$ is easy - simply factor out a variable from $n^{th}$-column/row and another variable from the $(n-1)^{th}$-column/row. These new columns/rows are identical.</p>\n\n<p>Inductive proofs neatly work with this Dodgson's recursive relation\n$$\\det Z^{0,0}(n)=\\frac{\\det Z^{1,1}(n-1)\\det Z^{1,1}(n-1)-\\det Z^{0,1}(n-1)\\det Z^{1,0}(n-1)}{\\det Z^{1,1}(n-2)}$$\nsatisfied by any matrix (so long as denominators do not vanish). Thus, it holds for $\\det M^{a,b}(n)$.\nSo, it remains to prove that the (explicit) formula on the RHS of (1) <em>does</em> satisfy the same equation. However, this is quite a routine simplification (preferably with symbolic sofwares). The proof follows. $\\square$</p>\n" } ]
Question: minimum-maximum entries matrix Answer 1: <p>Let us write $$a_r=\frac{x_{r+1}}{x_r}$$ for $r=1\cdots n$. </p> <p>We can then write the matrix $M(n)$ in the form $$\begin{pmatrix} 1 &amp; a_1 &amp; a_1a_2&amp; \cdots &amp; a_1a_2\cdots a_{n-1} \\ a_1 &amp; 1 &amp; a_2&amp; \cdots &amp; a_2\cdots a_{n-1}\\ \vdots &amp; \vdots &amp; \vdots &amp;\ddots &amp; \vdots\\ a_1a_2\cdots a_{n-1}&amp; a_2\cdots a_{n-1} &amp;\cdots &amp; a_{n-1} &amp; 1\end{pmatrix} $$ We do now Gauss elimination, and reduce $a_{n-1}$ times the $(n-1)$-th row from the $n$-th row. We then get in the last row $0, 0, \ldots (1-a_{n-1}^2)$. But this means that $det(M(n))=det(M(n-1))(1-a_{n-1}^2)$ and by induction $$det(M(n)) = \prod_{r=1}^{n-1} (1-a_r^2)=\prod_{r=1}^{n-1}(1-\frac{x_{r+1}^2}{x_r^2}).$$ By using inductively the Gauss elimination mentioned above, one can also get the inverse of $M(n)$.</p> Answer 2: <p>I wish to advertise the <a href="https://en.wikipedia.org/wiki/Dodgson_condensation" rel="nofollow noreferrer">Method of Condensation</a> in proving determinantal evaluation. The <strong>key</strong> is guess the answer, which is thanks to Ehud Meir. Then, generalize it a bit. Let $x_1, x_2,\dots$ be an infinite set of variables and modify the original matrix (by shifting variables) to $M^{a,b}(n)$ so that $$M_{i,j}^{a,b}(n):=\frac{x_{\max(i+a,j+b)}}{x_{\min(i+a,j+b)}}.$$ <em>Convention:</em> $M^{0,0}(n)=M(n)$. </p> <p><strong>Claim.</strong> If $a\neq b$ then $\det M^{a,b}(n)=0$, and if $a=b$ then $$\det M^{a,a}(n)=\prod_{r=2}^n\frac{x_{k-1+a}^2-x_{k+a}^2}{x_{k-1+a}^2}.\tag1$$ <strong>Proof.</strong> The case $a\neq b$ is easy - simply factor out a variable from $n^{th}$-column/row and another variable from the $(n-1)^{th}$-column/row. These new columns/rows are identical.</p> <p>Inductive proofs neatly work with this Dodgson's recursive relation $$\det Z^{0,0}(n)=\frac{\det Z^{1,1}(n-1)\det Z^{1,1}(n-1)-\det Z^{0,1}(n-1)\det Z^{1,0}(n-1)}{\det Z^{1,1}(n-2)}$$ satisfied by any matrix (so long as denominators do not vanish). Thus, it holds for $\det M^{a,b}(n)$. So, it remains to prove that the (explicit) formula on the RHS of (1) <em>does</em> satisfy the same equation. However, this is quite a routine simplification (preferably with symbolic sofwares). The proof follows. $\square$</p> Answer 3: <p>Let us write $$a_r=\frac{x_{r+1}}{x_r}$$ for $r=1\cdots n$. </p> <p>We can then write the matrix $M(n)$ in the form $$\begin{pmatrix} 1 &amp; a_1 &amp; a_1a_2&amp; \cdots &amp; a_1a_2\cdots a_{n-1} \\ a_1 &amp; 1 &amp; a_2&amp; \cdots &amp; a_2\cdots a_{n-1}\\ \vdots &amp; \vdots &amp; \vdots &amp;\ddots &amp; \vdots\\ a_1a_2\cdots a_{n-1}&amp; a_2\cdots a_{n-1} &amp;\cdots &amp; a_{n-1} &amp; 1\end{pmatrix} $$ We do now Gauss elimination, and reduce $a_{n-1}$ times the $(n-1)$-th row from the $n$-th row. We then get in the last row $0, 0, \ldots (1-a_{n-1}^2)$. But this means that $det(M(n))=det(M(n-1))(1-a_{n-1}^2)$ and by induction $$det(M(n)) = \prod_{r=1}^{n-1} (1-a_r^2)=\prod_{r=1}^{n-1}(1-\frac{x_{r+1}^2}{x_r^2}).$$ By using inductively the Gauss elimination mentioned above, one can also get the inverse of $M(n)$.</p> Answer 4: <p>I wish to advertise the <a href="https://en.wikipedia.org/wiki/Dodgson_condensation" rel="nofollow noreferrer">Method of Condensation</a> in proving determinantal evaluation. The <strong>key</strong> is guess the answer, which is thanks to Ehud Meir. Then, generalize it a bit. Let $x_1, x_2,\dots$ be an infinite set of variables and modify the original matrix (by shifting variables) to $M^{a,b}(n)$ so that $$M_{i,j}^{a,b}(n):=\frac{x_{\max(i+a,j+b)}}{x_{\min(i+a,j+b)}}.$$ <em>Convention:</em> $M^{0,0}(n)=M(n)$. </p> <p><strong>Claim.</strong> If $a\neq b$ then $\det M^{a,b}(n)=0$, and if $a=b$ then $$\det M^{a,a}(n)=\prod_{r=2}^n\frac{x_{k-1+a}^2-x_{k+a}^2}{x_{k-1+a}^2}.\tag1$$ <strong>Proof.</strong> The case $a\neq b$ is easy - simply factor out a variable from $n^{th}$-column/row and another variable from the $(n-1)^{th}$-column/row. These new columns/rows are identical.</p> <p>Inductive proofs neatly work with this Dodgson's recursive relation $$\det Z^{0,0}(n)=\frac{\det Z^{1,1}(n-1)\det Z^{1,1}(n-1)-\det Z^{0,1}(n-1)\det Z^{1,0}(n-1)}{\det Z^{1,1}(n-2)}$$ satisfied by any matrix (so long as denominators do not vanish). Thus, it holds for $\det M^{a,b}(n)$. So, it remains to prove that the (explicit) formula on the RHS of (1) <em>does</em> satisfy the same equation. However, this is quite a routine simplification (preferably with symbolic sofwares). The proof follows. $\square$</p>
Countable partitions of Cantor space mod meager
258,804
5
https://mathoverflow.net/questions/258804/countable-partitions-of-cantor-space-mod-meager
<p>Let $I$ be an index set. Given $A\subseteq I\times 2^\omega$ and $i\in I$, set $(A)_i = \{x\in 2^\omega: (i, x)\in A\}$. </p> <p>Now let $I\times 2^\omega = \bigcup_{n&lt;\omega} A_n$. How large must $I$ be to ensure that there are indices $i\neq j\in I$ and $n&lt; \omega$ with $(A_n)_i \cap (A_n)_j$ non-meager? In particular, is it consistent that there is a partition of $\omega_1\times 2^w$ into countably many pieces not satisfying the above?</p>
[ { "answer_id": 258807, "score": 6, "text": "<p><strong>The Borel case.</strong> I claim that every uncountable index set $I$ has your property, if the sets $A^i_n$ appearing in the covers are Borel. It follows that the answer to your follow-up particular question about $\\omega_1$ is negative for the Borel case. </p>\n\n<p>Suppose that $I$ is uncountable. Let me set things up a little differently than you did. What you\nreally have is $I$ many countable coverings of the reals. That\nis, for each $i\\in I$, we have a countable covering of the\nreals $\\bigcup_n A^i_n$. </p>\n\n<p>(So $\\bigcup_i (\\{i\\}\\times\nA^i_n)$ is what you call $A_n$, but I got confused thinking about it that way.)</p>\n\n<p>We know by the Baire category theorem that for any particular\n$i$, there is some non-meager $A^i_{n_i}$ for some $n_i$. Consider the map $i\\mapsto\nn_i$. Since $I$ is\nuncountable, this map cannot be injective, and in fact for\nsome $n$, it must be that uncountably many $i$ all select the same $n$. So we've got uncountably many $i$ for which $A^i_n$ is non-meager. Since in the Borel sets the meager ideal has the countable chain condition, it cannot be that these sets form an antichain modulo the meager ideal. So there must be $i\\neq j$ with $A^i_n\\cap A^j_n$ non-meager, and in fact uncountably many like that.</p>\n\n<p><strong>General case.</strong> In the general case, the argument shows that any set $I$ whose size is at least the chain condition of the meager ideal will have your property. Basically, we argue just as above. If $\\delta$ is the chain condition of the meager ideal, and we have $A^i_n$ with $\\bigcup_n A^i_n$ covering the reals, then we get $n_i$ constant for $\\delta$ many $i$, contradicting the chain condition, and so $A^i_n\\cap A^j_n$ is non-meager for some $i\\neq j$. </p>\n\n<p>Unfortunately, I don't know much about how big the chain condition of the meager ideal can be. There are some results mentioned in <a href=\"https://books.google.com/books?id=DLCyehuI0i0C&amp;pg=PA1061&amp;dq=meager+ideal+ccc&amp;hl=en&amp;sa=X&amp;ved=0ahUKEwjG_530h6rRAhXSZiYKHS1-ABsQ6AEIHDAA#v=onepage&amp;q=meager%20ideal%20ccc&amp;f=false\" rel=\"noreferrer\">Foreman's handbook article</a>. </p>\n" }, { "answer_id": 258851, "score": 2, "text": "<p>Let $\\kappa$ be least such that whenever $\\{A^i_n : n &lt; \\omega\\}$, for $i &lt; \\kappa$ are coverings of $2^{\\omega}$, there are $i &lt; j &lt; \\kappa$ and $n &lt; \\omega$ such that $A^i_n \\cap A^j_n$ is non meager.</p>\n\n<p>Claim: CH implies $\\kappa &gt; \\omega_1$.</p>\n\n<p>Proof: Using an Ulam matrix.</p>\n\n<p>Claim: Assume MA plus not CH. Then $\\kappa = \\omega_1$.</p>\n\n<p>Proof: Let $\\{A^i_n: n &lt; \\omega\\}$ be coverings of $2^{\\omega}$ for $i &lt; \\omega_1$. Towards a contradiction, assume that for every $i &lt; j &lt; \\omega_1$ and $n &lt; \\omega$, $A^i_n \\cap A^j_n$ is meager. Put $B^i_n = A^i_n \\setminus \\bigcup \\{A^j_n : j &lt; i\\}$ and $B^i = \\bigcup \\{B^i_n : n &lt; \\omega\\}$. Then each $B^i$ is comeager and so is $B = \\bigcap \\{B^i : i &lt; \\omega_1\\}$ (by MA plus not CH). But now if $x \\in B$, then $(\\forall i &lt; \\omega_1)(\\exists n &lt; \\omega)(x \\in B^i_n)$. Choose $i &lt; j &lt; \\omega_1$ and $n &lt; \\omega$ such that $x \\in B^i_n \\cap B^j_n$: Contradiction.</p>\n\n<p>So the lower bound $\\omega_1 \\leq \\kappa$ cannot be improved in ZFC. I am not sure about the optimal upper bound for $\\kappa$.</p>\n" }, { "answer_id": 258807, "score": 6, "text": "<p><strong>The Borel case.</strong> I claim that every uncountable index set $I$ has your property, if the sets $A^i_n$ appearing in the covers are Borel. It follows that the answer to your follow-up particular question about $\\omega_1$ is negative for the Borel case. </p>\n\n<p>Suppose that $I$ is uncountable. Let me set things up a little differently than you did. What you\nreally have is $I$ many countable coverings of the reals. That\nis, for each $i\\in I$, we have a countable covering of the\nreals $\\bigcup_n A^i_n$. </p>\n\n<p>(So $\\bigcup_i (\\{i\\}\\times\nA^i_n)$ is what you call $A_n$, but I got confused thinking about it that way.)</p>\n\n<p>We know by the Baire category theorem that for any particular\n$i$, there is some non-meager $A^i_{n_i}$ for some $n_i$. Consider the map $i\\mapsto\nn_i$. Since $I$ is\nuncountable, this map cannot be injective, and in fact for\nsome $n$, it must be that uncountably many $i$ all select the same $n$. So we've got uncountably many $i$ for which $A^i_n$ is non-meager. Since in the Borel sets the meager ideal has the countable chain condition, it cannot be that these sets form an antichain modulo the meager ideal. So there must be $i\\neq j$ with $A^i_n\\cap A^j_n$ non-meager, and in fact uncountably many like that.</p>\n\n<p><strong>General case.</strong> In the general case, the argument shows that any set $I$ whose size is at least the chain condition of the meager ideal will have your property. Basically, we argue just as above. If $\\delta$ is the chain condition of the meager ideal, and we have $A^i_n$ with $\\bigcup_n A^i_n$ covering the reals, then we get $n_i$ constant for $\\delta$ many $i$, contradicting the chain condition, and so $A^i_n\\cap A^j_n$ is non-meager for some $i\\neq j$. </p>\n\n<p>Unfortunately, I don't know much about how big the chain condition of the meager ideal can be. There are some results mentioned in <a href=\"https://books.google.com/books?id=DLCyehuI0i0C&amp;pg=PA1061&amp;dq=meager+ideal+ccc&amp;hl=en&amp;sa=X&amp;ved=0ahUKEwjG_530h6rRAhXSZiYKHS1-ABsQ6AEIHDAA#v=onepage&amp;q=meager%20ideal%20ccc&amp;f=false\" rel=\"noreferrer\">Foreman's handbook article</a>. </p>\n" }, { "answer_id": 258851, "score": 2, "text": "<p>Let $\\kappa$ be least such that whenever $\\{A^i_n : n &lt; \\omega\\}$, for $i &lt; \\kappa$ are coverings of $2^{\\omega}$, there are $i &lt; j &lt; \\kappa$ and $n &lt; \\omega$ such that $A^i_n \\cap A^j_n$ is non meager.</p>\n\n<p>Claim: CH implies $\\kappa &gt; \\omega_1$.</p>\n\n<p>Proof: Using an Ulam matrix.</p>\n\n<p>Claim: Assume MA plus not CH. Then $\\kappa = \\omega_1$.</p>\n\n<p>Proof: Let $\\{A^i_n: n &lt; \\omega\\}$ be coverings of $2^{\\omega}$ for $i &lt; \\omega_1$. Towards a contradiction, assume that for every $i &lt; j &lt; \\omega_1$ and $n &lt; \\omega$, $A^i_n \\cap A^j_n$ is meager. Put $B^i_n = A^i_n \\setminus \\bigcup \\{A^j_n : j &lt; i\\}$ and $B^i = \\bigcup \\{B^i_n : n &lt; \\omega\\}$. Then each $B^i$ is comeager and so is $B = \\bigcap \\{B^i : i &lt; \\omega_1\\}$ (by MA plus not CH). But now if $x \\in B$, then $(\\forall i &lt; \\omega_1)(\\exists n &lt; \\omega)(x \\in B^i_n)$. Choose $i &lt; j &lt; \\omega_1$ and $n &lt; \\omega$ such that $x \\in B^i_n \\cap B^j_n$: Contradiction.</p>\n\n<p>So the lower bound $\\omega_1 \\leq \\kappa$ cannot be improved in ZFC. I am not sure about the optimal upper bound for $\\kappa$.</p>\n" } ]
Question: Countable partitions of Cantor space mod meager Answer 1: <p><strong>The Borel case.</strong> I claim that every uncountable index set $I$ has your property, if the sets $A^i_n$ appearing in the covers are Borel. It follows that the answer to your follow-up particular question about $\omega_1$ is negative for the Borel case. </p> <p>Suppose that $I$ is uncountable. Let me set things up a little differently than you did. What you really have is $I$ many countable coverings of the reals. That is, for each $i\in I$, we have a countable covering of the reals $\bigcup_n A^i_n$. </p> <p>(So $\bigcup_i (\{i\}\times A^i_n)$ is what you call $A_n$, but I got confused thinking about it that way.)</p> <p>We know by the Baire category theorem that for any particular $i$, there is some non-meager $A^i_{n_i}$ for some $n_i$. Consider the map $i\mapsto n_i$. Since $I$ is uncountable, this map cannot be injective, and in fact for some $n$, it must be that uncountably many $i$ all select the same $n$. So we've got uncountably many $i$ for which $A^i_n$ is non-meager. Since in the Borel sets the meager ideal has the countable chain condition, it cannot be that these sets form an antichain modulo the meager ideal. So there must be $i\neq j$ with $A^i_n\cap A^j_n$ non-meager, and in fact uncountably many like that.</p> <p><strong>General case.</strong> In the general case, the argument shows that any set $I$ whose size is at least the chain condition of the meager ideal will have your property. Basically, we argue just as above. If $\delta$ is the chain condition of the meager ideal, and we have $A^i_n$ with $\bigcup_n A^i_n$ covering the reals, then we get $n_i$ constant for $\delta$ many $i$, contradicting the chain condition, and so $A^i_n\cap A^j_n$ is non-meager for some $i\neq j$. </p> <p>Unfortunately, I don't know much about how big the chain condition of the meager ideal can be. There are some results mentioned in <a href="https://books.google.com/books?id=DLCyehuI0i0C&amp;pg=PA1061&amp;dq=meager+ideal+ccc&amp;hl=en&amp;sa=X&amp;ved=0ahUKEwjG_530h6rRAhXSZiYKHS1-ABsQ6AEIHDAA#v=onepage&amp;q=meager%20ideal%20ccc&amp;f=false" rel="noreferrer">Foreman's handbook article</a>. </p> Answer 2: <p>Let $\kappa$ be least such that whenever $\{A^i_n : n &lt; \omega\}$, for $i &lt; \kappa$ are coverings of $2^{\omega}$, there are $i &lt; j &lt; \kappa$ and $n &lt; \omega$ such that $A^i_n \cap A^j_n$ is non meager.</p> <p>Claim: CH implies $\kappa &gt; \omega_1$.</p> <p>Proof: Using an Ulam matrix.</p> <p>Claim: Assume MA plus not CH. Then $\kappa = \omega_1$.</p> <p>Proof: Let $\{A^i_n: n &lt; \omega\}$ be coverings of $2^{\omega}$ for $i &lt; \omega_1$. Towards a contradiction, assume that for every $i &lt; j &lt; \omega_1$ and $n &lt; \omega$, $A^i_n \cap A^j_n$ is meager. Put $B^i_n = A^i_n \setminus \bigcup \{A^j_n : j &lt; i\}$ and $B^i = \bigcup \{B^i_n : n &lt; \omega\}$. Then each $B^i$ is comeager and so is $B = \bigcap \{B^i : i &lt; \omega_1\}$ (by MA plus not CH). But now if $x \in B$, then $(\forall i &lt; \omega_1)(\exists n &lt; \omega)(x \in B^i_n)$. Choose $i &lt; j &lt; \omega_1$ and $n &lt; \omega$ such that $x \in B^i_n \cap B^j_n$: Contradiction.</p> <p>So the lower bound $\omega_1 \leq \kappa$ cannot be improved in ZFC. I am not sure about the optimal upper bound for $\kappa$.</p> Answer 3: <p><strong>The Borel case.</strong> I claim that every uncountable index set $I$ has your property, if the sets $A^i_n$ appearing in the covers are Borel. It follows that the answer to your follow-up particular question about $\omega_1$ is negative for the Borel case. </p> <p>Suppose that $I$ is uncountable. Let me set things up a little differently than you did. What you really have is $I$ many countable coverings of the reals. That is, for each $i\in I$, we have a countable covering of the reals $\bigcup_n A^i_n$. </p> <p>(So $\bigcup_i (\{i\}\times A^i_n)$ is what you call $A_n$, but I got confused thinking about it that way.)</p> <p>We know by the Baire category theorem that for any particular $i$, there is some non-meager $A^i_{n_i}$ for some $n_i$. Consider the map $i\mapsto n_i$. Since $I$ is uncountable, this map cannot be injective, and in fact for some $n$, it must be that uncountably many $i$ all select the same $n$. So we've got uncountably many $i$ for which $A^i_n$ is non-meager. Since in the Borel sets the meager ideal has the countable chain condition, it cannot be that these sets form an antichain modulo the meager ideal. So there must be $i\neq j$ with $A^i_n\cap A^j_n$ non-meager, and in fact uncountably many like that.</p> <p><strong>General case.</strong> In the general case, the argument shows that any set $I$ whose size is at least the chain condition of the meager ideal will have your property. Basically, we argue just as above. If $\delta$ is the chain condition of the meager ideal, and we have $A^i_n$ with $\bigcup_n A^i_n$ covering the reals, then we get $n_i$ constant for $\delta$ many $i$, contradicting the chain condition, and so $A^i_n\cap A^j_n$ is non-meager for some $i\neq j$. </p> <p>Unfortunately, I don't know much about how big the chain condition of the meager ideal can be. There are some results mentioned in <a href="https://books.google.com/books?id=DLCyehuI0i0C&amp;pg=PA1061&amp;dq=meager+ideal+ccc&amp;hl=en&amp;sa=X&amp;ved=0ahUKEwjG_530h6rRAhXSZiYKHS1-ABsQ6AEIHDAA#v=onepage&amp;q=meager%20ideal%20ccc&amp;f=false" rel="noreferrer">Foreman's handbook article</a>. </p> Answer 4: <p>Let $\kappa$ be least such that whenever $\{A^i_n : n &lt; \omega\}$, for $i &lt; \kappa$ are coverings of $2^{\omega}$, there are $i &lt; j &lt; \kappa$ and $n &lt; \omega$ such that $A^i_n \cap A^j_n$ is non meager.</p> <p>Claim: CH implies $\kappa &gt; \omega_1$.</p> <p>Proof: Using an Ulam matrix.</p> <p>Claim: Assume MA plus not CH. Then $\kappa = \omega_1$.</p> <p>Proof: Let $\{A^i_n: n &lt; \omega\}$ be coverings of $2^{\omega}$ for $i &lt; \omega_1$. Towards a contradiction, assume that for every $i &lt; j &lt; \omega_1$ and $n &lt; \omega$, $A^i_n \cap A^j_n$ is meager. Put $B^i_n = A^i_n \setminus \bigcup \{A^j_n : j &lt; i\}$ and $B^i = \bigcup \{B^i_n : n &lt; \omega\}$. Then each $B^i$ is comeager and so is $B = \bigcap \{B^i : i &lt; \omega_1\}$ (by MA plus not CH). But now if $x \in B$, then $(\forall i &lt; \omega_1)(\exists n &lt; \omega)(x \in B^i_n)$. Choose $i &lt; j &lt; \omega_1$ and $n &lt; \omega$ such that $x \in B^i_n \cap B^j_n$: Contradiction.</p> <p>So the lower bound $\omega_1 \leq \kappa$ cannot be improved in ZFC. I am not sure about the optimal upper bound for $\kappa$.</p>
An interesting variant on the maximum independent set problem.
258,816
5
https://mathoverflow.net/questions/258816/an-interesting-variant-on-the-maximum-independent-set-problem
<p>Suppose i have a graph $G=(V,E)$ with $|V|=n$. Furthermore suppose i give you a maximum independent set $\mathcal{I}$ in $G$. Now suppose i obtain a new graph $G'$ from $G$ by removing a single vertex $v \in \mathcal{I}$ from $V$. Now of course $\alpha(G') \in \{|\mathcal{I}|,|\mathcal{I}|-1\}$ but can we say more? Is it an NP-hard to find a maximum independent set in $G'$? </p>
[ { "answer_id": 258868, "score": 3, "text": "<p>Suppose that $G$ is obtained from the complete bipartite graph with parts of size $n/2$ by adding some edges in each part. Then a maximum independent set lies in one part or the other, so determining $\\alpha(G')$ requires solving the maximum independent set problem in the part disjoint from $\\mathcal{I}$. So this is definitely NP-complete.</p>\n" }, { "answer_id": 258900, "score": 3, "text": "<p>Following the suggestion in <a href=\"https://mathoverflow.net/questions/258816/an-interesting-variant-on-the-maximum-independent-set-problem#comment638528_258816\">Ben Barber's comment</a>, any algorithm for solving your problem (of finding a maximum independent set in $G'$) can be used to find a maximum independent set in any finite graph.</p>\n\n<p>Let G be a graph of order $n.$ For $0\\le k\\le n$ let $G_k$ be a graph of order $n+k$ obtained by adding to $G$ an independent set $\\mathcal I_k$ of $k$ new vertices, and adding edges joining all the vertices in $\\mathcal I_k$ to all the vertices of $G.$</p>\n\n<p>Since $\\mathcal I_n$ is a maximum independent set in $G_n,$ we can use your algorithm to find a maximum independent set in $G_{n-1}.$ If the maximum independent set that we find in $G_{n-1}$ is $\\mathcal I_{n-1},$ we use the algorithm again to find a maximum independent set in $G_{n-2},$ and so on. Eventually we will find a maximum independent set of some $G_{k}$ which is contained in $V(G).$</p>\n" }, { "answer_id": 258868, "score": 3, "text": "<p>Suppose that $G$ is obtained from the complete bipartite graph with parts of size $n/2$ by adding some edges in each part. Then a maximum independent set lies in one part or the other, so determining $\\alpha(G')$ requires solving the maximum independent set problem in the part disjoint from $\\mathcal{I}$. So this is definitely NP-complete.</p>\n" }, { "answer_id": 258900, "score": 3, "text": "<p>Following the suggestion in <a href=\"https://mathoverflow.net/questions/258816/an-interesting-variant-on-the-maximum-independent-set-problem#comment638528_258816\">Ben Barber's comment</a>, any algorithm for solving your problem (of finding a maximum independent set in $G'$) can be used to find a maximum independent set in any finite graph.</p>\n\n<p>Let G be a graph of order $n.$ For $0\\le k\\le n$ let $G_k$ be a graph of order $n+k$ obtained by adding to $G$ an independent set $\\mathcal I_k$ of $k$ new vertices, and adding edges joining all the vertices in $\\mathcal I_k$ to all the vertices of $G.$</p>\n\n<p>Since $\\mathcal I_n$ is a maximum independent set in $G_n,$ we can use your algorithm to find a maximum independent set in $G_{n-1}.$ If the maximum independent set that we find in $G_{n-1}$ is $\\mathcal I_{n-1},$ we use the algorithm again to find a maximum independent set in $G_{n-2},$ and so on. Eventually we will find a maximum independent set of some $G_{k}$ which is contained in $V(G).$</p>\n" } ]
Question: An interesting variant on the maximum independent set problem. Answer 1: <p>Suppose that $G$ is obtained from the complete bipartite graph with parts of size $n/2$ by adding some edges in each part. Then a maximum independent set lies in one part or the other, so determining $\alpha(G')$ requires solving the maximum independent set problem in the part disjoint from $\mathcal{I}$. So this is definitely NP-complete.</p> Answer 2: <p>Following the suggestion in <a href="https://mathoverflow.net/questions/258816/an-interesting-variant-on-the-maximum-independent-set-problem#comment638528_258816">Ben Barber's comment</a>, any algorithm for solving your problem (of finding a maximum independent set in $G'$) can be used to find a maximum independent set in any finite graph.</p> <p>Let G be a graph of order $n.$ For $0\le k\le n$ let $G_k$ be a graph of order $n+k$ obtained by adding to $G$ an independent set $\mathcal I_k$ of $k$ new vertices, and adding edges joining all the vertices in $\mathcal I_k$ to all the vertices of $G.$</p> <p>Since $\mathcal I_n$ is a maximum independent set in $G_n,$ we can use your algorithm to find a maximum independent set in $G_{n-1}.$ If the maximum independent set that we find in $G_{n-1}$ is $\mathcal I_{n-1},$ we use the algorithm again to find a maximum independent set in $G_{n-2},$ and so on. Eventually we will find a maximum independent set of some $G_{k}$ which is contained in $V(G).$</p> Answer 3: <p>Suppose that $G$ is obtained from the complete bipartite graph with parts of size $n/2$ by adding some edges in each part. Then a maximum independent set lies in one part or the other, so determining $\alpha(G')$ requires solving the maximum independent set problem in the part disjoint from $\mathcal{I}$. So this is definitely NP-complete.</p> Answer 4: <p>Following the suggestion in <a href="https://mathoverflow.net/questions/258816/an-interesting-variant-on-the-maximum-independent-set-problem#comment638528_258816">Ben Barber's comment</a>, any algorithm for solving your problem (of finding a maximum independent set in $G'$) can be used to find a maximum independent set in any finite graph.</p> <p>Let G be a graph of order $n.$ For $0\le k\le n$ let $G_k$ be a graph of order $n+k$ obtained by adding to $G$ an independent set $\mathcal I_k$ of $k$ new vertices, and adding edges joining all the vertices in $\mathcal I_k$ to all the vertices of $G.$</p> <p>Since $\mathcal I_n$ is a maximum independent set in $G_n,$ we can use your algorithm to find a maximum independent set in $G_{n-1}.$ If the maximum independent set that we find in $G_{n-1}$ is $\mathcal I_{n-1},$ we use the algorithm again to find a maximum independent set in $G_{n-2},$ and so on. Eventually we will find a maximum independent set of some $G_{k}$ which is contained in $V(G).$</p>
Adjoint functor of power-set functor
258,934
5
https://mathoverflow.net/questions/258934/adjoint-functor-of-power-set-functor
<p>I want to ask a maybe stupid question. Power-set functor $P: \mathbf{Set}^{op} \rightarrow \mathbf{Set}$ is a well-known contravariant functor which sends each set to its powser set. A natural question is which functor is right(left) adjoint to it and I guess its oppostie functor $P^{op}: \mathbf{Set} \rightarrow \mathbf{Set}^{op}$ is right adjoint, is it true?</p>
[ { "answer_id": 258948, "score": 15, "text": "<p>Although the question has an easy answer (given in my comment under the question), it may be well to expand that answer by pointing out some related facts. </p>\n\n<p>So in the first place, we can say that in any cartesian closed category $\\mathbf{C}$, or even in any <a href=\"https://ncatlab.org/nlab/show/closed+monoidal+category\" rel=\"noreferrer\">symmetric monoidal closed</a> category meaning there is a family of adjunctions $- \\otimes A \\dashv [A, -]$ parametrized over objects $A$, every object $D$ induces an adjunction where $[-, D]^{op}: \\mathbf{C} \\to \\mathbf{C}^{op}$ is left adjoint to $[-, D]: \\mathbf{C}^{op} \\to \\mathbf{C}$. For there are natural bijections </p>\n\n<p>$$\\hom(A, [B, D]) \\cong \\hom(A \\otimes B, D) \\cong \\hom(B \\otimes A, D) \\cong \\hom(B, [A, D])$$ </p>\n\n<p>where the middle isomorphism uses symmetry. This may be rewritten as a natural bijection </p>\n\n<p>$$\\hom_{\\mathbf{C}^{op}}([B, D], A) \\cong \\hom_{\\mathbf{C}}(B, [A, D])$$ </p>\n\n<p>which expresses the adjunction $[-, D]^{op} \\dashv [-, D]$. </p>\n\n<p>A deeper fact is that in the special case $\\mathbf{C} = Set$ and $D = 2 = \\{0, 1\\}$, the functor </p>\n\n<p>$$P = [-, 2]: Set^{op} \\to Set$$ </p>\n\n<p>is <a href=\"https://ncatlab.org/nlab/show/monadic+functor\" rel=\"noreferrer\"><em>monadic</em></a>. That is to say: $Set^{op}$ is equivalent to a category of algebras $\\mathbf{A}$ of an algebraic theory (identifying $P$ with the underlying set-functor $U: \\mathbf{A} \\to Set$). In this case, $\\mathbf{A}$ is the category of <a href=\"https://ncatlab.org/nlab/show/complete+Boolean+algebra#cabas\" rel=\"noreferrer\">complete atomic Boolean algebras</a> (CABAs), or equivalently the category of <a href=\"https://ncatlab.org/nlab/show/completely+distributive+lattice#completely_distributive_boolean_algebras\" rel=\"noreferrer\">completely distributive complete Boolean algebras</a>. This is saying in particular that every CABA is a power set Boolean algebra $P(A)$, and conversely. (Contrast with the case of complete Boolean algebras, where in fact the underlying set-functor has no left adjoint.) </p>\n\n<p>One of the big theorems of elementary topos theory (due to Robert Par&eacute;) is the corresponding generalization, that if $\\Omega$ is the subobject classifier of a topos $\\mathbf{E}$, then </p>\n\n<p>$$P(-) = [-, \\Omega]: \\mathbf{E}^{op} \\to \\mathbf{E}$$ </p>\n\n<p>is also monadic. Since monadic functors reflect limits, this implies that finite limits exist in $\\mathbf{E}^{op}$ since they exist in $\\mathbf{E}$, and therefore a topos is finitely cocomplete. </p>\n" }, { "answer_id": 258939, "score": -1, "text": "<p>It does not have a left adjoint, because it is essentially the forgetful functor from complete and atomic boolean algebras to the category of sets.</p>\n\n<p>See <a href=\"https://ncatlab.org/nlab/show/complete+Boolean+algebra\" rel=\"nofollow noreferrer\">https://ncatlab.org/nlab/show/complete+Boolean+algebra</a>.</p>\n" }, { "answer_id": 258948, "score": 15, "text": "<p>Although the question has an easy answer (given in my comment under the question), it may be well to expand that answer by pointing out some related facts. </p>\n\n<p>So in the first place, we can say that in any cartesian closed category $\\mathbf{C}$, or even in any <a href=\"https://ncatlab.org/nlab/show/closed+monoidal+category\" rel=\"noreferrer\">symmetric monoidal closed</a> category meaning there is a family of adjunctions $- \\otimes A \\dashv [A, -]$ parametrized over objects $A$, every object $D$ induces an adjunction where $[-, D]^{op}: \\mathbf{C} \\to \\mathbf{C}^{op}$ is left adjoint to $[-, D]: \\mathbf{C}^{op} \\to \\mathbf{C}$. For there are natural bijections </p>\n\n<p>$$\\hom(A, [B, D]) \\cong \\hom(A \\otimes B, D) \\cong \\hom(B \\otimes A, D) \\cong \\hom(B, [A, D])$$ </p>\n\n<p>where the middle isomorphism uses symmetry. This may be rewritten as a natural bijection </p>\n\n<p>$$\\hom_{\\mathbf{C}^{op}}([B, D], A) \\cong \\hom_{\\mathbf{C}}(B, [A, D])$$ </p>\n\n<p>which expresses the adjunction $[-, D]^{op} \\dashv [-, D]$. </p>\n\n<p>A deeper fact is that in the special case $\\mathbf{C} = Set$ and $D = 2 = \\{0, 1\\}$, the functor </p>\n\n<p>$$P = [-, 2]: Set^{op} \\to Set$$ </p>\n\n<p>is <a href=\"https://ncatlab.org/nlab/show/monadic+functor\" rel=\"noreferrer\"><em>monadic</em></a>. That is to say: $Set^{op}$ is equivalent to a category of algebras $\\mathbf{A}$ of an algebraic theory (identifying $P$ with the underlying set-functor $U: \\mathbf{A} \\to Set$). In this case, $\\mathbf{A}$ is the category of <a href=\"https://ncatlab.org/nlab/show/complete+Boolean+algebra#cabas\" rel=\"noreferrer\">complete atomic Boolean algebras</a> (CABAs), or equivalently the category of <a href=\"https://ncatlab.org/nlab/show/completely+distributive+lattice#completely_distributive_boolean_algebras\" rel=\"noreferrer\">completely distributive complete Boolean algebras</a>. This is saying in particular that every CABA is a power set Boolean algebra $P(A)$, and conversely. (Contrast with the case of complete Boolean algebras, where in fact the underlying set-functor has no left adjoint.) </p>\n\n<p>One of the big theorems of elementary topos theory (due to Robert Par&eacute;) is the corresponding generalization, that if $\\Omega$ is the subobject classifier of a topos $\\mathbf{E}$, then </p>\n\n<p>$$P(-) = [-, \\Omega]: \\mathbf{E}^{op} \\to \\mathbf{E}$$ </p>\n\n<p>is also monadic. Since monadic functors reflect limits, this implies that finite limits exist in $\\mathbf{E}^{op}$ since they exist in $\\mathbf{E}$, and therefore a topos is finitely cocomplete. </p>\n" }, { "answer_id": 258939, "score": -1, "text": "<p>It does not have a left adjoint, because it is essentially the forgetful functor from complete and atomic boolean algebras to the category of sets.</p>\n\n<p>See <a href=\"https://ncatlab.org/nlab/show/complete+Boolean+algebra\" rel=\"nofollow noreferrer\">https://ncatlab.org/nlab/show/complete+Boolean+algebra</a>.</p>\n" } ]
Question: Adjoint functor of power-set functor Answer 1: <p>Although the question has an easy answer (given in my comment under the question), it may be well to expand that answer by pointing out some related facts. </p> <p>So in the first place, we can say that in any cartesian closed category $\mathbf{C}$, or even in any <a href="https://ncatlab.org/nlab/show/closed+monoidal+category" rel="noreferrer">symmetric monoidal closed</a> category meaning there is a family of adjunctions $- \otimes A \dashv [A, -]$ parametrized over objects $A$, every object $D$ induces an adjunction where $[-, D]^{op}: \mathbf{C} \to \mathbf{C}^{op}$ is left adjoint to $[-, D]: \mathbf{C}^{op} \to \mathbf{C}$. For there are natural bijections </p> <p>$$\hom(A, [B, D]) \cong \hom(A \otimes B, D) \cong \hom(B \otimes A, D) \cong \hom(B, [A, D])$$ </p> <p>where the middle isomorphism uses symmetry. This may be rewritten as a natural bijection </p> <p>$$\hom_{\mathbf{C}^{op}}([B, D], A) \cong \hom_{\mathbf{C}}(B, [A, D])$$ </p> <p>which expresses the adjunction $[-, D]^{op} \dashv [-, D]$. </p> <p>A deeper fact is that in the special case $\mathbf{C} = Set$ and $D = 2 = \{0, 1\}$, the functor </p> <p>$$P = [-, 2]: Set^{op} \to Set$$ </p> <p>is <a href="https://ncatlab.org/nlab/show/monadic+functor" rel="noreferrer"><em>monadic</em></a>. That is to say: $Set^{op}$ is equivalent to a category of algebras $\mathbf{A}$ of an algebraic theory (identifying $P$ with the underlying set-functor $U: \mathbf{A} \to Set$). In this case, $\mathbf{A}$ is the category of <a href="https://ncatlab.org/nlab/show/complete+Boolean+algebra#cabas" rel="noreferrer">complete atomic Boolean algebras</a> (CABAs), or equivalently the category of <a href="https://ncatlab.org/nlab/show/completely+distributive+lattice#completely_distributive_boolean_algebras" rel="noreferrer">completely distributive complete Boolean algebras</a>. This is saying in particular that every CABA is a power set Boolean algebra $P(A)$, and conversely. (Contrast with the case of complete Boolean algebras, where in fact the underlying set-functor has no left adjoint.) </p> <p>One of the big theorems of elementary topos theory (due to Robert Par&eacute;) is the corresponding generalization, that if $\Omega$ is the subobject classifier of a topos $\mathbf{E}$, then </p> <p>$$P(-) = [-, \Omega]: \mathbf{E}^{op} \to \mathbf{E}$$ </p> <p>is also monadic. Since monadic functors reflect limits, this implies that finite limits exist in $\mathbf{E}^{op}$ since they exist in $\mathbf{E}$, and therefore a topos is finitely cocomplete. </p> Answer 2: <p>It does not have a left adjoint, because it is essentially the forgetful functor from complete and atomic boolean algebras to the category of sets.</p> <p>See <a href="https://ncatlab.org/nlab/show/complete+Boolean+algebra" rel="nofollow noreferrer">https://ncatlab.org/nlab/show/complete+Boolean+algebra</a>.</p> Answer 3: <p>Although the question has an easy answer (given in my comment under the question), it may be well to expand that answer by pointing out some related facts. </p> <p>So in the first place, we can say that in any cartesian closed category $\mathbf{C}$, or even in any <a href="https://ncatlab.org/nlab/show/closed+monoidal+category" rel="noreferrer">symmetric monoidal closed</a> category meaning there is a family of adjunctions $- \otimes A \dashv [A, -]$ parametrized over objects $A$, every object $D$ induces an adjunction where $[-, D]^{op}: \mathbf{C} \to \mathbf{C}^{op}$ is left adjoint to $[-, D]: \mathbf{C}^{op} \to \mathbf{C}$. For there are natural bijections </p> <p>$$\hom(A, [B, D]) \cong \hom(A \otimes B, D) \cong \hom(B \otimes A, D) \cong \hom(B, [A, D])$$ </p> <p>where the middle isomorphism uses symmetry. This may be rewritten as a natural bijection </p> <p>$$\hom_{\mathbf{C}^{op}}([B, D], A) \cong \hom_{\mathbf{C}}(B, [A, D])$$ </p> <p>which expresses the adjunction $[-, D]^{op} \dashv [-, D]$. </p> <p>A deeper fact is that in the special case $\mathbf{C} = Set$ and $D = 2 = \{0, 1\}$, the functor </p> <p>$$P = [-, 2]: Set^{op} \to Set$$ </p> <p>is <a href="https://ncatlab.org/nlab/show/monadic+functor" rel="noreferrer"><em>monadic</em></a>. That is to say: $Set^{op}$ is equivalent to a category of algebras $\mathbf{A}$ of an algebraic theory (identifying $P$ with the underlying set-functor $U: \mathbf{A} \to Set$). In this case, $\mathbf{A}$ is the category of <a href="https://ncatlab.org/nlab/show/complete+Boolean+algebra#cabas" rel="noreferrer">complete atomic Boolean algebras</a> (CABAs), or equivalently the category of <a href="https://ncatlab.org/nlab/show/completely+distributive+lattice#completely_distributive_boolean_algebras" rel="noreferrer">completely distributive complete Boolean algebras</a>. This is saying in particular that every CABA is a power set Boolean algebra $P(A)$, and conversely. (Contrast with the case of complete Boolean algebras, where in fact the underlying set-functor has no left adjoint.) </p> <p>One of the big theorems of elementary topos theory (due to Robert Par&eacute;) is the corresponding generalization, that if $\Omega$ is the subobject classifier of a topos $\mathbf{E}$, then </p> <p>$$P(-) = [-, \Omega]: \mathbf{E}^{op} \to \mathbf{E}$$ </p> <p>is also monadic. Since monadic functors reflect limits, this implies that finite limits exist in $\mathbf{E}^{op}$ since they exist in $\mathbf{E}$, and therefore a topos is finitely cocomplete. </p> Answer 4: <p>It does not have a left adjoint, because it is essentially the forgetful functor from complete and atomic boolean algebras to the category of sets.</p> <p>See <a href="https://ncatlab.org/nlab/show/complete+Boolean+algebra" rel="nofollow noreferrer">https://ncatlab.org/nlab/show/complete+Boolean+algebra</a>.</p>
Is there a Bailey–Borwein–Plouffe (BBP) formula for e?
259,053
5
https://mathoverflow.net/questions/259053/is-there-a-bailey-borwein-plouffe-bbp-formula-for-e
<p>I recently used Bailey–Borwein–Plouffe formula to implement a <a href="https://www.browserling.com/tools/pi-digits" rel="nofollow noreferrer">π digit generator</a>. Now I also want to implement an <code>e</code> digit generator, for the <a href="https://en.wikipedia.org/wiki/E_(mathematical_constant)" rel="nofollow noreferrer">Euler number</a>. </p> <p>I've been searching and reading papers but it looks like there is no BBP formula for <code>e</code>. Is there one? Does anyone know if there has been recent progress on this?</p>
[ { "answer_id": 259059, "score": 6, "text": "<p>I wish I had 50+ reputation so I could comment : I visited your website and I saw that you are not actually using what the BBP is famous for namely extracting the $n$-th digit in base-16. Instead you are giving <em>all</em> the digits in base 10. If that is what you want for $e$ then the power series $\\sum_{k \\geq 0} 1/k!$ already does the job sufficiently well. </p>\n" }, { "answer_id": 259062, "score": 1, "text": "<p>What you are actually looking for, is the <a href=\"https://en.wikipedia.org/wiki/Spigot_algorithm\" rel=\"nofollow noreferrer\">Spigot Algorithm</a> which is also known as \"droplet algorithm\".<br>\nThat algorithm has been described a few years ago in the german \"Spektrum der Wissenschaft\" and should thus also have been described in the \"Scientific American\".</p>\n\n<p>Especially the algorithm for $e$ is known and an implementation in R can be found <a href=\"https://www.rdocumentation.org/packages/numbers/versions/0.6-1/topics/dropletPi\" rel=\"nofollow noreferrer\">here</a>.</p>\n" }, { "answer_id": 259059, "score": 6, "text": "<p>I wish I had 50+ reputation so I could comment : I visited your website and I saw that you are not actually using what the BBP is famous for namely extracting the $n$-th digit in base-16. Instead you are giving <em>all</em> the digits in base 10. If that is what you want for $e$ then the power series $\\sum_{k \\geq 0} 1/k!$ already does the job sufficiently well. </p>\n" }, { "answer_id": 259062, "score": 1, "text": "<p>What you are actually looking for, is the <a href=\"https://en.wikipedia.org/wiki/Spigot_algorithm\" rel=\"nofollow noreferrer\">Spigot Algorithm</a> which is also known as \"droplet algorithm\".<br>\nThat algorithm has been described a few years ago in the german \"Spektrum der Wissenschaft\" and should thus also have been described in the \"Scientific American\".</p>\n\n<p>Especially the algorithm for $e$ is known and an implementation in R can be found <a href=\"https://www.rdocumentation.org/packages/numbers/versions/0.6-1/topics/dropletPi\" rel=\"nofollow noreferrer\">here</a>.</p>\n" } ]
Question: Is there a Bailey–Borwein–Plouffe (BBP) formula for e? Answer 1: <p>I wish I had 50+ reputation so I could comment : I visited your website and I saw that you are not actually using what the BBP is famous for namely extracting the $n$-th digit in base-16. Instead you are giving <em>all</em> the digits in base 10. If that is what you want for $e$ then the power series $\sum_{k \geq 0} 1/k!$ already does the job sufficiently well. </p> Answer 2: <p>What you are actually looking for, is the <a href="https://en.wikipedia.org/wiki/Spigot_algorithm" rel="nofollow noreferrer">Spigot Algorithm</a> which is also known as "droplet algorithm".<br> That algorithm has been described a few years ago in the german "Spektrum der Wissenschaft" and should thus also have been described in the "Scientific American".</p> <p>Especially the algorithm for $e$ is known and an implementation in R can be found <a href="https://www.rdocumentation.org/packages/numbers/versions/0.6-1/topics/dropletPi" rel="nofollow noreferrer">here</a>.</p> Answer 3: <p>I wish I had 50+ reputation so I could comment : I visited your website and I saw that you are not actually using what the BBP is famous for namely extracting the $n$-th digit in base-16. Instead you are giving <em>all</em> the digits in base 10. If that is what you want for $e$ then the power series $\sum_{k \geq 0} 1/k!$ already does the job sufficiently well. </p> Answer 4: <p>What you are actually looking for, is the <a href="https://en.wikipedia.org/wiki/Spigot_algorithm" rel="nofollow noreferrer">Spigot Algorithm</a> which is also known as "droplet algorithm".<br> That algorithm has been described a few years ago in the german "Spektrum der Wissenschaft" and should thus also have been described in the "Scientific American".</p> <p>Especially the algorithm for $e$ is known and an implementation in R can be found <a href="https://www.rdocumentation.org/packages/numbers/versions/0.6-1/topics/dropletPi" rel="nofollow noreferrer">here</a>.</p>
On a dual of Kaplansky&#39;s $2^{nd}$ conjecture: admissible algebras?
259,058
5
https://mathoverflow.net/questions/259058/on-a-dual-of-kaplanskys-2nd-conjecture-admissible-algebras
<p><a href="http://www.mathematik.uni-muenchen.de/~sommerh/Publikationen/KaplConjwww/KaplConjwww_en.php" rel="nofollow noreferrer">Kaplansky's second conjecture</a> (on Hopf algebras) deals with "<em>admissible</em>" coalgebras: He calls a coalgebra admissible, if there is an algebra structure making it a Hopf algebra. The conjecture states that: </p> <blockquote> <p>a coalgebra $C$ is admissible if and only if any finite subset of $C$ lies in a finite dimensional admissible subcoalgebra. </p> </blockquote> <p>As far as I know there are counterexamples refuting the conjecture. (the first one was stated by Larson if I remember correctly). </p> <p>Inspired by the above, let us lay the following definition of the notion of <strong>admissible algebra</strong>: </p> <blockquote> <p><strong>Definition:</strong> An algebra $A$ will be called admissible if there is a coalgebra structure on $A$ and a suitable linear map $S:A\to A$, such that all these data (the algebra, the coalgebra and the linear map $S$) constitute a Hopf algebra with antipode $S$. </p> </blockquote> <p>Now, my question is whether there is some criterion (necessary, sufficient or both) regarding to when an algebra is admissible (or non-admissible) in the sense of the above definition. In other words: </p> <ul> <li>Given an algebra $A$, under which conditions can a coalgebra structure be found on $A$ such that it will be turned into a Hopf algebra? </li> <li>When would such a coalgebra structure be unique? </li> <li>What could be examples of a <strong>non-admissible algebras</strong>, in the sense of the above definition? </li> </ul>
[ { "answer_id": 259952, "score": 3, "text": "<p>A couple of examples of <strong>non-admissible algebras</strong>: </p>\n\n<p>$\\bullet$ It is known (see <a href=\"https://books.google.gr/books?id=pBJ6sbPHA0IC&amp;printsec=frontcover&amp;hl=el#v=onepage&amp;q&amp;f=false\" rel=\"nofollow noreferrer\">here</a>, p.222) that a semisimple Hopf algebra is also separable (as an algebra). Consequently, <strong>a semisimple but non-separable</strong> algebra has to be non-admissible.<br>\n(An example of such an algebra could be the following: Consider a field $K$ and $L$ an inseparable extension. Then $L$ is a semisimple $K$-algebra but the algebra $L\\otimes L$ is no longer semisimple). </p>\n\n<p>$\\bullet$ On the other hand, another example (outside the semisimple algebras) might be the case of <strong>Weyl algebras and their deformations</strong>. As far as I know, there is no known Hopf structure for the family of Weyl algebras. Furthermore, in an old paper of T.D. Palev, it is argued that it is impossible to even define a comultiplication on Weyl algebras. See: <a href=\"https://arxiv.org/abs/hep-th/9307032\" rel=\"nofollow noreferrer\">Is it possible to extend the deformed Weyl algebra $W_q(n)$ to a Hopf algebra?</a> (the discussion of p.3-5 refers to the undeformed case). </p>\n\n<p><strong>Edit:</strong><br>\n$\\bullet$ I have recently found an interesting result leading to some other classes of examples of non-admissible algebras:</p>\n\n<blockquote>\n <p><strong>Proposition:</strong> Let $A$ be a connected graded algebra and let $x,y$ be non-zero elements in $A$ such that $xy=qyx$ for some $q\\in k\\backslash \\{1\\}$. Then there is no Hopf algebra structure on $A$. </p>\n</blockquote>\n\n<p>Thus, $A$ is non-admissible in the sense of the OP's definition. This is proposition 2.7 from <a href=\"https://arxiv.org/abs/1601.06687v1\" rel=\"nofollow noreferrer\">Connected (graded) hopf algebras</a>. One of the authors appears to be the P.Gilmartin (who has provided the other answer above). So I am adding this reference and the following examples for the shake of completeness. As a consequence of the above proposition we get the following families of examples of non-admissible algebras (i am copying Corollary 2.8 of the formerly mentioned paper): </p>\n\n<ul>\n<li>Sklyanin algebras of any dimension</li>\n<li>skew-polynomial rings $k_{p_{i,j}}[x_1,x_2,...,x_n]$ (except for the case $p_{i,j}=1$ for all $i,j$) </li>\n<li>Quantum matrix algebras $\\mathcal{O}_q(M_{n\\times n})$ for $q\\neq 1$</li>\n<li>non-commutative Koszul Artin-Schelter regular algebras of dimension $\\leq 4$</li>\n</ul>\n\n<p>$\\bullet$ Using the above proposition (for $q=-1$), we can conclude that the <strong>fermionic algebra</strong> or the <strong>algebra of anticommutation relations</strong>, generated as an algebra by the fermion creation-annihilation operators $f_i^+, f_i^-$ for $i=1,2,...$ modulo the following relations\n$$\n\\{f_i^-,f_j^+\\}=\\delta_{ij}, \\ \\ \\{f_i^-,f_j^-\\}=0, \\ \\ \\{f_i^+,f_j^+\\}=0\n$$\n(where $\\{x,y\\}=xy+yx$) for all $i,j=1,2,...$ is another example of non-admissible algebra i.e. of an algebra which does not admit a Hopf algebra structure. This might be of particular interest to mathematical physicists since this algebra describes the spin-$\\frac{1}{2}$ particles.<br>\n(This last example is due to me and is not contained in the paper cited above. So if there is any disagreement ... put the blame on me!)</p>\n" }, { "answer_id": 264544, "score": 3, "text": "<p>In regards to your first question: I don't believe there is a way of telling in general whether or not an arbitrary algebra $A$ admits a Hopf structure. There are, however, in certain settings, some conditions an algebra $A$ must satisfy in order to be a Hopf algebra. \n\\</p>\n\n<p>For example, if $A$ is a Hopf algebra over an algebraically closed field of characteristic zero, its abelianisation $A_{ab}=A/\\langle [A, A] \\rangle$ is a commutative Hopf algebra (with structure inherited from $A$ - this part is true for any Hopf algebra over any field), and, by the Nullstellensatz, is the coordinate ring of a smooth algebraic group. In particular, $A_{ab}$ is a commutative domain. Thus, it must be that any algebra $B$ with an abelianisation $B_{ab}$ which is NOT a domain cannot be an admissible algebra in the above sense. An example of such an algebra would be $k[x,y]/\\langle [x,y] = y^2 \\rangle$. This example also answers your third question.</p>\n\n<p>In regards to your second question about uniqueness: in general an algebra can admit several distinct Hopf structures. Take, for example, $H = k[x,y,z]$. Its Hopf structures are in 1-1 correspondence with 3-d unipotent groups, of which there are two: the abelian group and the 3-d Heisenberg group. These correspond to a Hopf structure generated by primitive elements in the former case, and a Hopf structure generated by two primitives ($x$ and $y$ say) and a non-primitive, $z$ say, such that $\\Delta(z) = 1\\otimes z + x\\otimes y + z\\otimes 1$.</p>\n\n<p>You can read a bit more about the examples I've given here:</p>\n\n<p><a href=\"https://arxiv.org/abs/1506.02427\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1506.02427</a></p>\n" }, { "answer_id": 308852, "score": 2, "text": "<p>Note that a Hopf algebra is augmented. Let $A$ be an augmented algebra over a commutative ring $k$. A <strong>necessary condition</strong> for $A$ having a Hopf algebra structure is that the cohomology ring \n$$H^\\ast(A,k) := Ext_A^\\ast(k,k)$$\nis graded commutative under Yoneda composition. </p>\n\n<p>In fact, the coproduct of a Hopf algebra (or, more generally, of a bialgebra) induces a graded comutative cup product on cohomology and one can show that Yoneda composition and cup product agree on $Ext_A^\\ast(k,k)$. </p>\n\n<p>When I have time, I'll look for an example there this criterion applies. </p>\n" }, { "answer_id": 259952, "score": 3, "text": "<p>A couple of examples of <strong>non-admissible algebras</strong>: </p>\n\n<p>$\\bullet$ It is known (see <a href=\"https://books.google.gr/books?id=pBJ6sbPHA0IC&amp;printsec=frontcover&amp;hl=el#v=onepage&amp;q&amp;f=false\" rel=\"nofollow noreferrer\">here</a>, p.222) that a semisimple Hopf algebra is also separable (as an algebra). Consequently, <strong>a semisimple but non-separable</strong> algebra has to be non-admissible.<br>\n(An example of such an algebra could be the following: Consider a field $K$ and $L$ an inseparable extension. Then $L$ is a semisimple $K$-algebra but the algebra $L\\otimes L$ is no longer semisimple). </p>\n\n<p>$\\bullet$ On the other hand, another example (outside the semisimple algebras) might be the case of <strong>Weyl algebras and their deformations</strong>. As far as I know, there is no known Hopf structure for the family of Weyl algebras. Furthermore, in an old paper of T.D. Palev, it is argued that it is impossible to even define a comultiplication on Weyl algebras. See: <a href=\"https://arxiv.org/abs/hep-th/9307032\" rel=\"nofollow noreferrer\">Is it possible to extend the deformed Weyl algebra $W_q(n)$ to a Hopf algebra?</a> (the discussion of p.3-5 refers to the undeformed case). </p>\n\n<p><strong>Edit:</strong><br>\n$\\bullet$ I have recently found an interesting result leading to some other classes of examples of non-admissible algebras:</p>\n\n<blockquote>\n <p><strong>Proposition:</strong> Let $A$ be a connected graded algebra and let $x,y$ be non-zero elements in $A$ such that $xy=qyx$ for some $q\\in k\\backslash \\{1\\}$. Then there is no Hopf algebra structure on $A$. </p>\n</blockquote>\n\n<p>Thus, $A$ is non-admissible in the sense of the OP's definition. This is proposition 2.7 from <a href=\"https://arxiv.org/abs/1601.06687v1\" rel=\"nofollow noreferrer\">Connected (graded) hopf algebras</a>. One of the authors appears to be the P.Gilmartin (who has provided the other answer above). So I am adding this reference and the following examples for the shake of completeness. As a consequence of the above proposition we get the following families of examples of non-admissible algebras (i am copying Corollary 2.8 of the formerly mentioned paper): </p>\n\n<ul>\n<li>Sklyanin algebras of any dimension</li>\n<li>skew-polynomial rings $k_{p_{i,j}}[x_1,x_2,...,x_n]$ (except for the case $p_{i,j}=1$ for all $i,j$) </li>\n<li>Quantum matrix algebras $\\mathcal{O}_q(M_{n\\times n})$ for $q\\neq 1$</li>\n<li>non-commutative Koszul Artin-Schelter regular algebras of dimension $\\leq 4$</li>\n</ul>\n\n<p>$\\bullet$ Using the above proposition (for $q=-1$), we can conclude that the <strong>fermionic algebra</strong> or the <strong>algebra of anticommutation relations</strong>, generated as an algebra by the fermion creation-annihilation operators $f_i^+, f_i^-$ for $i=1,2,...$ modulo the following relations\n$$\n\\{f_i^-,f_j^+\\}=\\delta_{ij}, \\ \\ \\{f_i^-,f_j^-\\}=0, \\ \\ \\{f_i^+,f_j^+\\}=0\n$$\n(where $\\{x,y\\}=xy+yx$) for all $i,j=1,2,...$ is another example of non-admissible algebra i.e. of an algebra which does not admit a Hopf algebra structure. This might be of particular interest to mathematical physicists since this algebra describes the spin-$\\frac{1}{2}$ particles.<br>\n(This last example is due to me and is not contained in the paper cited above. So if there is any disagreement ... put the blame on me!)</p>\n" }, { "answer_id": 264544, "score": 3, "text": "<p>In regards to your first question: I don't believe there is a way of telling in general whether or not an arbitrary algebra $A$ admits a Hopf structure. There are, however, in certain settings, some conditions an algebra $A$ must satisfy in order to be a Hopf algebra. \n\\</p>\n\n<p>For example, if $A$ is a Hopf algebra over an algebraically closed field of characteristic zero, its abelianisation $A_{ab}=A/\\langle [A, A] \\rangle$ is a commutative Hopf algebra (with structure inherited from $A$ - this part is true for any Hopf algebra over any field), and, by the Nullstellensatz, is the coordinate ring of a smooth algebraic group. In particular, $A_{ab}$ is a commutative domain. Thus, it must be that any algebra $B$ with an abelianisation $B_{ab}$ which is NOT a domain cannot be an admissible algebra in the above sense. An example of such an algebra would be $k[x,y]/\\langle [x,y] = y^2 \\rangle$. This example also answers your third question.</p>\n\n<p>In regards to your second question about uniqueness: in general an algebra can admit several distinct Hopf structures. Take, for example, $H = k[x,y,z]$. Its Hopf structures are in 1-1 correspondence with 3-d unipotent groups, of which there are two: the abelian group and the 3-d Heisenberg group. These correspond to a Hopf structure generated by primitive elements in the former case, and a Hopf structure generated by two primitives ($x$ and $y$ say) and a non-primitive, $z$ say, such that $\\Delta(z) = 1\\otimes z + x\\otimes y + z\\otimes 1$.</p>\n\n<p>You can read a bit more about the examples I've given here:</p>\n\n<p><a href=\"https://arxiv.org/abs/1506.02427\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1506.02427</a></p>\n" }, { "answer_id": 308852, "score": 2, "text": "<p>Note that a Hopf algebra is augmented. Let $A$ be an augmented algebra over a commutative ring $k$. A <strong>necessary condition</strong> for $A$ having a Hopf algebra structure is that the cohomology ring \n$$H^\\ast(A,k) := Ext_A^\\ast(k,k)$$\nis graded commutative under Yoneda composition. </p>\n\n<p>In fact, the coproduct of a Hopf algebra (or, more generally, of a bialgebra) induces a graded comutative cup product on cohomology and one can show that Yoneda composition and cup product agree on $Ext_A^\\ast(k,k)$. </p>\n\n<p>When I have time, I'll look for an example there this criterion applies. </p>\n" } ]
Question: On a dual of Kaplansky&#39;s $2^{nd}$ conjecture: admissible algebras? Answer 1: <p>A couple of examples of <strong>non-admissible algebras</strong>: </p> <p>$\bullet$ It is known (see <a href="https://books.google.gr/books?id=pBJ6sbPHA0IC&amp;printsec=frontcover&amp;hl=el#v=onepage&amp;q&amp;f=false" rel="nofollow noreferrer">here</a>, p.222) that a semisimple Hopf algebra is also separable (as an algebra). Consequently, <strong>a semisimple but non-separable</strong> algebra has to be non-admissible.<br> (An example of such an algebra could be the following: Consider a field $K$ and $L$ an inseparable extension. Then $L$ is a semisimple $K$-algebra but the algebra $L\otimes L$ is no longer semisimple). </p> <p>$\bullet$ On the other hand, another example (outside the semisimple algebras) might be the case of <strong>Weyl algebras and their deformations</strong>. As far as I know, there is no known Hopf structure for the family of Weyl algebras. Furthermore, in an old paper of T.D. Palev, it is argued that it is impossible to even define a comultiplication on Weyl algebras. See: <a href="https://arxiv.org/abs/hep-th/9307032" rel="nofollow noreferrer">Is it possible to extend the deformed Weyl algebra $W_q(n)$ to a Hopf algebra?</a> (the discussion of p.3-5 refers to the undeformed case). </p> <p><strong>Edit:</strong><br> $\bullet$ I have recently found an interesting result leading to some other classes of examples of non-admissible algebras:</p> <blockquote> <p><strong>Proposition:</strong> Let $A$ be a connected graded algebra and let $x,y$ be non-zero elements in $A$ such that $xy=qyx$ for some $q\in k\backslash \{1\}$. Then there is no Hopf algebra structure on $A$. </p> </blockquote> <p>Thus, $A$ is non-admissible in the sense of the OP's definition. This is proposition 2.7 from <a href="https://arxiv.org/abs/1601.06687v1" rel="nofollow noreferrer">Connected (graded) hopf algebras</a>. One of the authors appears to be the P.Gilmartin (who has provided the other answer above). So I am adding this reference and the following examples for the shake of completeness. As a consequence of the above proposition we get the following families of examples of non-admissible algebras (i am copying Corollary 2.8 of the formerly mentioned paper): </p> <ul> <li>Sklyanin algebras of any dimension</li> <li>skew-polynomial rings $k_{p_{i,j}}[x_1,x_2,...,x_n]$ (except for the case $p_{i,j}=1$ for all $i,j$) </li> <li>Quantum matrix algebras $\mathcal{O}_q(M_{n\times n})$ for $q\neq 1$</li> <li>non-commutative Koszul Artin-Schelter regular algebras of dimension $\leq 4$</li> </ul> <p>$\bullet$ Using the above proposition (for $q=-1$), we can conclude that the <strong>fermionic algebra</strong> or the <strong>algebra of anticommutation relations</strong>, generated as an algebra by the fermion creation-annihilation operators $f_i^+, f_i^-$ for $i=1,2,...$ modulo the following relations $$ \{f_i^-,f_j^+\}=\delta_{ij}, \ \ \{f_i^-,f_j^-\}=0, \ \ \{f_i^+,f_j^+\}=0 $$ (where $\{x,y\}=xy+yx$) for all $i,j=1,2,...$ is another example of non-admissible algebra i.e. of an algebra which does not admit a Hopf algebra structure. This might be of particular interest to mathematical physicists since this algebra describes the spin-$\frac{1}{2}$ particles.<br> (This last example is due to me and is not contained in the paper cited above. So if there is any disagreement ... put the blame on me!)</p> Answer 2: <p>In regards to your first question: I don't believe there is a way of telling in general whether or not an arbitrary algebra $A$ admits a Hopf structure. There are, however, in certain settings, some conditions an algebra $A$ must satisfy in order to be a Hopf algebra. \</p> <p>For example, if $A$ is a Hopf algebra over an algebraically closed field of characteristic zero, its abelianisation $A_{ab}=A/\langle [A, A] \rangle$ is a commutative Hopf algebra (with structure inherited from $A$ - this part is true for any Hopf algebra over any field), and, by the Nullstellensatz, is the coordinate ring of a smooth algebraic group. In particular, $A_{ab}$ is a commutative domain. Thus, it must be that any algebra $B$ with an abelianisation $B_{ab}$ which is NOT a domain cannot be an admissible algebra in the above sense. An example of such an algebra would be $k[x,y]/\langle [x,y] = y^2 \rangle$. This example also answers your third question.</p> <p>In regards to your second question about uniqueness: in general an algebra can admit several distinct Hopf structures. Take, for example, $H = k[x,y,z]$. Its Hopf structures are in 1-1 correspondence with 3-d unipotent groups, of which there are two: the abelian group and the 3-d Heisenberg group. These correspond to a Hopf structure generated by primitive elements in the former case, and a Hopf structure generated by two primitives ($x$ and $y$ say) and a non-primitive, $z$ say, such that $\Delta(z) = 1\otimes z + x\otimes y + z\otimes 1$.</p> <p>You can read a bit more about the examples I've given here:</p> <p><a href="https://arxiv.org/abs/1506.02427" rel="nofollow noreferrer">https://arxiv.org/abs/1506.02427</a></p> Answer 3: <p>Note that a Hopf algebra is augmented. Let $A$ be an augmented algebra over a commutative ring $k$. A <strong>necessary condition</strong> for $A$ having a Hopf algebra structure is that the cohomology ring $$H^\ast(A,k) := Ext_A^\ast(k,k)$$ is graded commutative under Yoneda composition. </p> <p>In fact, the coproduct of a Hopf algebra (or, more generally, of a bialgebra) induces a graded comutative cup product on cohomology and one can show that Yoneda composition and cup product agree on $Ext_A^\ast(k,k)$. </p> <p>When I have time, I'll look for an example there this criterion applies. </p> Answer 4: <p>A couple of examples of <strong>non-admissible algebras</strong>: </p> <p>$\bullet$ It is known (see <a href="https://books.google.gr/books?id=pBJ6sbPHA0IC&amp;printsec=frontcover&amp;hl=el#v=onepage&amp;q&amp;f=false" rel="nofollow noreferrer">here</a>, p.222) that a semisimple Hopf algebra is also separable (as an algebra). Consequently, <strong>a semisimple but non-separable</strong> algebra has to be non-admissible.<br> (An example of such an algebra could be the following: Consider a field $K$ and $L$ an inseparable extension. Then $L$ is a semisimple $K$-algebra but the algebra $L\otimes L$ is no longer semisimple). </p> <p>$\bullet$ On the other hand, another example (outside the semisimple algebras) might be the case of <strong>Weyl algebras and their deformations</strong>. As far as I know, there is no known Hopf structure for the family of Weyl algebras. Furthermore, in an old paper of T.D. Palev, it is argued that it is impossible to even define a comultiplication on Weyl algebras. See: <a href="https://arxiv.org/abs/hep-th/9307032" rel="nofollow noreferrer">Is it possible to extend the deformed Weyl algebra $W_q(n)$ to a Hopf algebra?</a> (the discussion of p.3-5 refers to the undeformed case). </p> <p><strong>Edit:</strong><br> $\bullet$ I have recently found an interesting result leading to some other classes of examples of non-admissible algebras:</p> <blockquote> <p><strong>Proposition:</strong> Let $A$ be a connected graded algebra and let $x,y$ be non-zero elements in $A$ such that $xy=qyx$ for some $q\in k\backslash \{1\}$. Then there is no Hopf algebra structure on $A$. </p> </blockquote> <p>Thus, $A$ is non-admissible in the sense of the OP's definition. This is proposition 2.7 from <a href="https://arxiv.org/abs/1601.06687v1" rel="nofollow noreferrer">Connected (graded) hopf algebras</a>. One of the authors appears to be the P.Gilmartin (who has provided the other answer above). So I am adding this reference and the following examples for the shake of completeness. As a consequence of the above proposition we get the following families of examples of non-admissible algebras (i am copying Corollary 2.8 of the formerly mentioned paper): </p> <ul> <li>Sklyanin algebras of any dimension</li> <li>skew-polynomial rings $k_{p_{i,j}}[x_1,x_2,...,x_n]$ (except for the case $p_{i,j}=1$ for all $i,j$) </li> <li>Quantum matrix algebras $\mathcal{O}_q(M_{n\times n})$ for $q\neq 1$</li> <li>non-commutative Koszul Artin-Schelter regular algebras of dimension $\leq 4$</li> </ul> <p>$\bullet$ Using the above proposition (for $q=-1$), we can conclude that the <strong>fermionic algebra</strong> or the <strong>algebra of anticommutation relations</strong>, generated as an algebra by the fermion creation-annihilation operators $f_i^+, f_i^-$ for $i=1,2,...$ modulo the following relations $$ \{f_i^-,f_j^+\}=\delta_{ij}, \ \ \{f_i^-,f_j^-\}=0, \ \ \{f_i^+,f_j^+\}=0 $$ (where $\{x,y\}=xy+yx$) for all $i,j=1,2,...$ is another example of non-admissible algebra i.e. of an algebra which does not admit a Hopf algebra structure. This might be of particular interest to mathematical physicists since this algebra describes the spin-$\frac{1}{2}$ particles.<br> (This last example is due to me and is not contained in the paper cited above. So if there is any disagreement ... put the blame on me!)</p> Answer 5: <p>In regards to your first question: I don't believe there is a way of telling in general whether or not an arbitrary algebra $A$ admits a Hopf structure. There are, however, in certain settings, some conditions an algebra $A$ must satisfy in order to be a Hopf algebra. \</p> <p>For example, if $A$ is a Hopf algebra over an algebraically closed field of characteristic zero, its abelianisation $A_{ab}=A/\langle [A, A] \rangle$ is a commutative Hopf algebra (with structure inherited from $A$ - this part is true for any Hopf algebra over any field), and, by the Nullstellensatz, is the coordinate ring of a smooth algebraic group. In particular, $A_{ab}$ is a commutative domain. Thus, it must be that any algebra $B$ with an abelianisation $B_{ab}$ which is NOT a domain cannot be an admissible algebra in the above sense. An example of such an algebra would be $k[x,y]/\langle [x,y] = y^2 \rangle$. This example also answers your third question.</p> <p>In regards to your second question about uniqueness: in general an algebra can admit several distinct Hopf structures. Take, for example, $H = k[x,y,z]$. Its Hopf structures are in 1-1 correspondence with 3-d unipotent groups, of which there are two: the abelian group and the 3-d Heisenberg group. These correspond to a Hopf structure generated by primitive elements in the former case, and a Hopf structure generated by two primitives ($x$ and $y$ say) and a non-primitive, $z$ say, such that $\Delta(z) = 1\otimes z + x\otimes y + z\otimes 1$.</p> <p>You can read a bit more about the examples I've given here:</p> <p><a href="https://arxiv.org/abs/1506.02427" rel="nofollow noreferrer">https://arxiv.org/abs/1506.02427</a></p> Answer 6: <p>Note that a Hopf algebra is augmented. Let $A$ be an augmented algebra over a commutative ring $k$. A <strong>necessary condition</strong> for $A$ having a Hopf algebra structure is that the cohomology ring $$H^\ast(A,k) := Ext_A^\ast(k,k)$$ is graded commutative under Yoneda composition. </p> <p>In fact, the coproduct of a Hopf algebra (or, more generally, of a bialgebra) induces a graded comutative cup product on cohomology and one can show that Yoneda composition and cup product agree on $Ext_A^\ast(k,k)$. </p> <p>When I have time, I'll look for an example there this criterion applies. </p>
Distance between two metric spaces
259,253
5
https://mathoverflow.net/questions/259253/distance-between-two-metric-spaces
<p>I am given two metric spaces as two arrays of the same size. Each one is supposed to represent distance between vertices on a mesh in R^3. The meshes are assumed to have the same number of vertices and the correspondence betweeen the vertices is also given. Is there a way to find the a meaningful distance between these two matrices (other than the trivial ones)? I would like the distance be invariant under scaling : in other words if the first mesh's is just a scaling of the other mesh then I want the distance between the two corresponding metric spaces to be zero (even though the distances in the matrices have also been scaled).</p> <p>I know that Hausdorff distance measures the distance between two sets in the same metric space but note here that my question is a little different so that notion is not exactly useful. </p> <p>Thanks.</p>
[ { "answer_id": 292819, "score": 3, "text": "<p>Firstly, if you just have pairwise distances rather than coordinates of points in $\\mathbb{R}^3$, then you can attempt to embed the mesh into $\\mathbb{R}^3$ using Isomap. Under certain conditions, such as if you have a <a href=\"https://en.wikipedia.org/wiki/Cauchy%27s_theorem_(geometry)\" rel=\"nofollow noreferrer\">triangulated convex polyhedron</a>, the isometric embedding into $\\mathbb{R}^3$ is unique up to isometries of $\\mathbb{R}^3$.</p>\n\n<p>After you embed your meshes $X, Y$ as finite labelled subsets of $\\mathbb{R}^3$, the problem of finding the optimal scaling and isometry which minimises the error $\\sum_i \\lVert X_i - Y_i \\rVert_2^2$ is called <a href=\"https://en.wikipedia.org/wiki/Procrustes_analysis\" rel=\"nofollow noreferrer\">Procrustes analysis</a>. After applying this isometry, you get a canonical (root-mean-square) distance between the sets $X, Y$.</p>\n\n<p>The way that Procrustes analysis usually proceeds is to subtract the mean, then scale to have unit variance, and then use singular value decomposition to find the optimal orthogonal matrix transforming $X$ to match $Y$ as closely as possible. The last part is called the <a href=\"https://en.wikipedia.org/wiki/Kabsch_algorithm\" rel=\"nofollow noreferrer\">Kabsch algorithm</a>.</p>\n" }, { "answer_id": 395747, "score": 0, "text": "<p>Persistent Homology can be used for this purpose. Given two metric spaces <span class=\"math-container\">$X$</span> and <span class=\"math-container\">$Y$</span>, then one can measure the distance between their persistence diagrams <span class=\"math-container\">$PD_i(X)$</span> and <span class=\"math-container\">$PD_i(Y)$</span> where <span class=\"math-container\">$i \\geq 0$</span>. There are many distances that can be defined on the space of persistence diagrams such as the <a href=\"http://gudhi.gforge.inria.fr/doc/latest/group__bottleneck__distance.html#:%7E:text=The%20bottleneck%20distance%20is%20the,Persistence%20Diagrams%22%20%5B34%5D.\" rel=\"nofollow noreferrer\">bottleneck distance</a> of the <a href=\"https://gudhi.inria.fr/python/3.1.0.rc1/wasserstein_distance_user.html\" rel=\"nofollow noreferrer\">Wasserstein distance</a>. The distance between the persistence diagram is much more efficient than the Hausdorff distance because usually the persistence diagram contains less data than the original spaces.</p>\n" }, { "answer_id": 292819, "score": 3, "text": "<p>Firstly, if you just have pairwise distances rather than coordinates of points in $\\mathbb{R}^3$, then you can attempt to embed the mesh into $\\mathbb{R}^3$ using Isomap. Under certain conditions, such as if you have a <a href=\"https://en.wikipedia.org/wiki/Cauchy%27s_theorem_(geometry)\" rel=\"nofollow noreferrer\">triangulated convex polyhedron</a>, the isometric embedding into $\\mathbb{R}^3$ is unique up to isometries of $\\mathbb{R}^3$.</p>\n\n<p>After you embed your meshes $X, Y$ as finite labelled subsets of $\\mathbb{R}^3$, the problem of finding the optimal scaling and isometry which minimises the error $\\sum_i \\lVert X_i - Y_i \\rVert_2^2$ is called <a href=\"https://en.wikipedia.org/wiki/Procrustes_analysis\" rel=\"nofollow noreferrer\">Procrustes analysis</a>. After applying this isometry, you get a canonical (root-mean-square) distance between the sets $X, Y$.</p>\n\n<p>The way that Procrustes analysis usually proceeds is to subtract the mean, then scale to have unit variance, and then use singular value decomposition to find the optimal orthogonal matrix transforming $X$ to match $Y$ as closely as possible. The last part is called the <a href=\"https://en.wikipedia.org/wiki/Kabsch_algorithm\" rel=\"nofollow noreferrer\">Kabsch algorithm</a>.</p>\n" }, { "answer_id": 395747, "score": 0, "text": "<p>Persistent Homology can be used for this purpose. Given two metric spaces <span class=\"math-container\">$X$</span> and <span class=\"math-container\">$Y$</span>, then one can measure the distance between their persistence diagrams <span class=\"math-container\">$PD_i(X)$</span> and <span class=\"math-container\">$PD_i(Y)$</span> where <span class=\"math-container\">$i \\geq 0$</span>. There are many distances that can be defined on the space of persistence diagrams such as the <a href=\"http://gudhi.gforge.inria.fr/doc/latest/group__bottleneck__distance.html#:%7E:text=The%20bottleneck%20distance%20is%20the,Persistence%20Diagrams%22%20%5B34%5D.\" rel=\"nofollow noreferrer\">bottleneck distance</a> of the <a href=\"https://gudhi.inria.fr/python/3.1.0.rc1/wasserstein_distance_user.html\" rel=\"nofollow noreferrer\">Wasserstein distance</a>. The distance between the persistence diagram is much more efficient than the Hausdorff distance because usually the persistence diagram contains less data than the original spaces.</p>\n" } ]
Question: Distance between two metric spaces Answer 1: <p>Firstly, if you just have pairwise distances rather than coordinates of points in $\mathbb{R}^3$, then you can attempt to embed the mesh into $\mathbb{R}^3$ using Isomap. Under certain conditions, such as if you have a <a href="https://en.wikipedia.org/wiki/Cauchy%27s_theorem_(geometry)" rel="nofollow noreferrer">triangulated convex polyhedron</a>, the isometric embedding into $\mathbb{R}^3$ is unique up to isometries of $\mathbb{R}^3$.</p> <p>After you embed your meshes $X, Y$ as finite labelled subsets of $\mathbb{R}^3$, the problem of finding the optimal scaling and isometry which minimises the error $\sum_i \lVert X_i - Y_i \rVert_2^2$ is called <a href="https://en.wikipedia.org/wiki/Procrustes_analysis" rel="nofollow noreferrer">Procrustes analysis</a>. After applying this isometry, you get a canonical (root-mean-square) distance between the sets $X, Y$.</p> <p>The way that Procrustes analysis usually proceeds is to subtract the mean, then scale to have unit variance, and then use singular value decomposition to find the optimal orthogonal matrix transforming $X$ to match $Y$ as closely as possible. The last part is called the <a href="https://en.wikipedia.org/wiki/Kabsch_algorithm" rel="nofollow noreferrer">Kabsch algorithm</a>.</p> Answer 2: <p>Persistent Homology can be used for this purpose. Given two metric spaces <span class="math-container">$X$</span> and <span class="math-container">$Y$</span>, then one can measure the distance between their persistence diagrams <span class="math-container">$PD_i(X)$</span> and <span class="math-container">$PD_i(Y)$</span> where <span class="math-container">$i \geq 0$</span>. There are many distances that can be defined on the space of persistence diagrams such as the <a href="http://gudhi.gforge.inria.fr/doc/latest/group__bottleneck__distance.html#:%7E:text=The%20bottleneck%20distance%20is%20the,Persistence%20Diagrams%22%20%5B34%5D." rel="nofollow noreferrer">bottleneck distance</a> of the <a href="https://gudhi.inria.fr/python/3.1.0.rc1/wasserstein_distance_user.html" rel="nofollow noreferrer">Wasserstein distance</a>. The distance between the persistence diagram is much more efficient than the Hausdorff distance because usually the persistence diagram contains less data than the original spaces.</p> Answer 3: <p>Firstly, if you just have pairwise distances rather than coordinates of points in $\mathbb{R}^3$, then you can attempt to embed the mesh into $\mathbb{R}^3$ using Isomap. Under certain conditions, such as if you have a <a href="https://en.wikipedia.org/wiki/Cauchy%27s_theorem_(geometry)" rel="nofollow noreferrer">triangulated convex polyhedron</a>, the isometric embedding into $\mathbb{R}^3$ is unique up to isometries of $\mathbb{R}^3$.</p> <p>After you embed your meshes $X, Y$ as finite labelled subsets of $\mathbb{R}^3$, the problem of finding the optimal scaling and isometry which minimises the error $\sum_i \lVert X_i - Y_i \rVert_2^2$ is called <a href="https://en.wikipedia.org/wiki/Procrustes_analysis" rel="nofollow noreferrer">Procrustes analysis</a>. After applying this isometry, you get a canonical (root-mean-square) distance between the sets $X, Y$.</p> <p>The way that Procrustes analysis usually proceeds is to subtract the mean, then scale to have unit variance, and then use singular value decomposition to find the optimal orthogonal matrix transforming $X$ to match $Y$ as closely as possible. The last part is called the <a href="https://en.wikipedia.org/wiki/Kabsch_algorithm" rel="nofollow noreferrer">Kabsch algorithm</a>.</p> Answer 4: <p>Persistent Homology can be used for this purpose. Given two metric spaces <span class="math-container">$X$</span> and <span class="math-container">$Y$</span>, then one can measure the distance between their persistence diagrams <span class="math-container">$PD_i(X)$</span> and <span class="math-container">$PD_i(Y)$</span> where <span class="math-container">$i \geq 0$</span>. There are many distances that can be defined on the space of persistence diagrams such as the <a href="http://gudhi.gforge.inria.fr/doc/latest/group__bottleneck__distance.html#:%7E:text=The%20bottleneck%20distance%20is%20the,Persistence%20Diagrams%22%20%5B34%5D." rel="nofollow noreferrer">bottleneck distance</a> of the <a href="https://gudhi.inria.fr/python/3.1.0.rc1/wasserstein_distance_user.html" rel="nofollow noreferrer">Wasserstein distance</a>. The distance between the persistence diagram is much more efficient than the Hausdorff distance because usually the persistence diagram contains less data than the original spaces.</p>
Expressing properties of graded algebras in terms of the $\mathbb{G}_m$action
259,535
5
https://mathoverflow.net/questions/259535/expressing-properties-of-graded-algebras-in-terms-of-the-mathbbg-maction
<p>Let us fix a base ring $k$. The category of $\mathbb{Z}$-graded $k$-algebras is equivalent to the category of $\mathbb{G}_m$ equivariant affine $k$-schemes. The following 2 properties often come up when constructing the $Proj$ of a graded ring: Let $A$ be a graded $k$-algebra.</p> <ol> <li>The zeroth graded piece is the base ring $A_0 =k$ </li> <li>The algebra $A_\bullet$ is positively graded. i.e. $A_j=0$ for all $j&lt;0$.</li> <li>The canonical multiplication homomorphism $Sym^\bullet A_1 \to A_\bullet$ is surjective.</li> </ol> <p>I'm trying to find equivalent characterization of the above properties which don't use the aforementioned equivalence. In other words I want the above 3 properties expressed in terms of properties of $\mathbb{G}_m$-equivariant schemes (orbits, stabilizers, fixed points etc.).</p>
[ { "answer_id": 259538, "score": 2, "text": "<p>Let $V$ be an affine variety with a $\\mathbb{G}_m$-action. Then you should think of $A_0$ as a good model for $V/\\mathbb{G}_m$. So being connected says something like $\\mathbb{G}_m$ acts transitively- at least from the point of view of functions. Being positively graded is saying something like the action of $\\mathbb{G}_m$ extending to an action of the monoid scheme $\\mathbb{A}^1$ (under multiplication). The last condition is more involved: The module $A_1$ is the set of $\\mathbb{G}_m$-equivariant functions to $\\mathbb{A}^1$ where we use the scaling action on $\\mathbb{A}^1$. (So it's kind of like \"linear\" functions on $V$). This is a vector space over $k$, call its <em>dual</em> $W$ and view it as a variety. There's a natural equivariant map $V \\to W$ given by evaluation (that's why we took the dual of the vector space of functions). We're asking that this equivariant map be an embedding. So, while $V$ is not a vector space with the standard linear action of $\\mathbb{G}_m$, it can be embedded in one in some natural way.</p>\n" }, { "answer_id": 259543, "score": 2, "text": "<p>Expanding on Dylan Wilson's answer, condition (2) is indeed just a continuation of $\\mathbb{G}_m$-action to an $\\mathbb{A}^1$-action, since $\\mathcal{O}(\\mathbb{A}^1) \\hookrightarrow \\mathcal{O}(\\mathbb{G}_m): \\mathbb{k}[t] \\hookrightarrow \\mathbb{k}[t, t^{-1}]$. Condition (1) is trickier, since the action of $\\mathbb{G}_m$ on $V$ will have inseparable orbits --- see the action on $\\mathbb{A}^n$. Thus we can't think of it as transitivity (not of the group action at least, the monoid $\\mathbb{A}^1$ action is transitive in a sense that any two points can be mapped to a same point). Instead we should look at it together with condition (2), which implies that the action of $\\mathbb{G}_m$ has fixed points: for any $x \\in V$ the point $0\\cdot x \\in V$ will be fixed. Here $0 \\in \\mathbb{A}^1$ and its action exists by (2). Condition (1) then reduces to the statement that the $\\mathbb{G}_m$-action has a <em>unique</em> fixed point. On the level of functions it is given by a graded morphism $\\mathcal{O}(V) \\to \\mathbb k$ as graded algebras. Condition (3) means that the action is <em>free</em> if we remove that single fixed point. We can see it because the action on $A_1^*$ is a free action apart from $0 \\in A_1^*$ and the embedding $V \\hookrightarrow A_1^*$ is $\\mathbb{G}_m$-equivariant, as a morphism of graded algebras.</p>\n\n<p>The simplest example where you can see all those statements is the obvious action on $\\mathbb{A}^n$. On the other hand, consider the obvious action of $\\mathbb{G}_m$ on itself. It corresponds to the natural grading on $\\mathbb{k}[t,t^{-1}]$. The reasons (2) fails here is obvious. Since there is no morphism $\\mathbb{k}[t,t^{-1}] \\to \\mathbb{k}$, the action has no fixed points. In this case (1) indeed reduces to transitivity. Instead of (3) we have that $Sym(A_1 \\oplus A_{-1}) \\to A_\\bullet$ is surjective. This also implies that the action is free, for the same reasons as above.</p>\n" }, { "answer_id": 259538, "score": 2, "text": "<p>Let $V$ be an affine variety with a $\\mathbb{G}_m$-action. Then you should think of $A_0$ as a good model for $V/\\mathbb{G}_m$. So being connected says something like $\\mathbb{G}_m$ acts transitively- at least from the point of view of functions. Being positively graded is saying something like the action of $\\mathbb{G}_m$ extending to an action of the monoid scheme $\\mathbb{A}^1$ (under multiplication). The last condition is more involved: The module $A_1$ is the set of $\\mathbb{G}_m$-equivariant functions to $\\mathbb{A}^1$ where we use the scaling action on $\\mathbb{A}^1$. (So it's kind of like \"linear\" functions on $V$). This is a vector space over $k$, call its <em>dual</em> $W$ and view it as a variety. There's a natural equivariant map $V \\to W$ given by evaluation (that's why we took the dual of the vector space of functions). We're asking that this equivariant map be an embedding. So, while $V$ is not a vector space with the standard linear action of $\\mathbb{G}_m$, it can be embedded in one in some natural way.</p>\n" }, { "answer_id": 259543, "score": 2, "text": "<p>Expanding on Dylan Wilson's answer, condition (2) is indeed just a continuation of $\\mathbb{G}_m$-action to an $\\mathbb{A}^1$-action, since $\\mathcal{O}(\\mathbb{A}^1) \\hookrightarrow \\mathcal{O}(\\mathbb{G}_m): \\mathbb{k}[t] \\hookrightarrow \\mathbb{k}[t, t^{-1}]$. Condition (1) is trickier, since the action of $\\mathbb{G}_m$ on $V$ will have inseparable orbits --- see the action on $\\mathbb{A}^n$. Thus we can't think of it as transitivity (not of the group action at least, the monoid $\\mathbb{A}^1$ action is transitive in a sense that any two points can be mapped to a same point). Instead we should look at it together with condition (2), which implies that the action of $\\mathbb{G}_m$ has fixed points: for any $x \\in V$ the point $0\\cdot x \\in V$ will be fixed. Here $0 \\in \\mathbb{A}^1$ and its action exists by (2). Condition (1) then reduces to the statement that the $\\mathbb{G}_m$-action has a <em>unique</em> fixed point. On the level of functions it is given by a graded morphism $\\mathcal{O}(V) \\to \\mathbb k$ as graded algebras. Condition (3) means that the action is <em>free</em> if we remove that single fixed point. We can see it because the action on $A_1^*$ is a free action apart from $0 \\in A_1^*$ and the embedding $V \\hookrightarrow A_1^*$ is $\\mathbb{G}_m$-equivariant, as a morphism of graded algebras.</p>\n\n<p>The simplest example where you can see all those statements is the obvious action on $\\mathbb{A}^n$. On the other hand, consider the obvious action of $\\mathbb{G}_m$ on itself. It corresponds to the natural grading on $\\mathbb{k}[t,t^{-1}]$. The reasons (2) fails here is obvious. Since there is no morphism $\\mathbb{k}[t,t^{-1}] \\to \\mathbb{k}$, the action has no fixed points. In this case (1) indeed reduces to transitivity. Instead of (3) we have that $Sym(A_1 \\oplus A_{-1}) \\to A_\\bullet$ is surjective. This also implies that the action is free, for the same reasons as above.</p>\n" } ]
Question: Expressing properties of graded algebras in terms of the $\mathbb{G}_m$action Answer 1: <p>Let $V$ be an affine variety with a $\mathbb{G}_m$-action. Then you should think of $A_0$ as a good model for $V/\mathbb{G}_m$. So being connected says something like $\mathbb{G}_m$ acts transitively- at least from the point of view of functions. Being positively graded is saying something like the action of $\mathbb{G}_m$ extending to an action of the monoid scheme $\mathbb{A}^1$ (under multiplication). The last condition is more involved: The module $A_1$ is the set of $\mathbb{G}_m$-equivariant functions to $\mathbb{A}^1$ where we use the scaling action on $\mathbb{A}^1$. (So it's kind of like "linear" functions on $V$). This is a vector space over $k$, call its <em>dual</em> $W$ and view it as a variety. There's a natural equivariant map $V \to W$ given by evaluation (that's why we took the dual of the vector space of functions). We're asking that this equivariant map be an embedding. So, while $V$ is not a vector space with the standard linear action of $\mathbb{G}_m$, it can be embedded in one in some natural way.</p> Answer 2: <p>Expanding on Dylan Wilson's answer, condition (2) is indeed just a continuation of $\mathbb{G}_m$-action to an $\mathbb{A}^1$-action, since $\mathcal{O}(\mathbb{A}^1) \hookrightarrow \mathcal{O}(\mathbb{G}_m): \mathbb{k}[t] \hookrightarrow \mathbb{k}[t, t^{-1}]$. Condition (1) is trickier, since the action of $\mathbb{G}_m$ on $V$ will have inseparable orbits --- see the action on $\mathbb{A}^n$. Thus we can't think of it as transitivity (not of the group action at least, the monoid $\mathbb{A}^1$ action is transitive in a sense that any two points can be mapped to a same point). Instead we should look at it together with condition (2), which implies that the action of $\mathbb{G}_m$ has fixed points: for any $x \in V$ the point $0\cdot x \in V$ will be fixed. Here $0 \in \mathbb{A}^1$ and its action exists by (2). Condition (1) then reduces to the statement that the $\mathbb{G}_m$-action has a <em>unique</em> fixed point. On the level of functions it is given by a graded morphism $\mathcal{O}(V) \to \mathbb k$ as graded algebras. Condition (3) means that the action is <em>free</em> if we remove that single fixed point. We can see it because the action on $A_1^*$ is a free action apart from $0 \in A_1^*$ and the embedding $V \hookrightarrow A_1^*$ is $\mathbb{G}_m$-equivariant, as a morphism of graded algebras.</p> <p>The simplest example where you can see all those statements is the obvious action on $\mathbb{A}^n$. On the other hand, consider the obvious action of $\mathbb{G}_m$ on itself. It corresponds to the natural grading on $\mathbb{k}[t,t^{-1}]$. The reasons (2) fails here is obvious. Since there is no morphism $\mathbb{k}[t,t^{-1}] \to \mathbb{k}$, the action has no fixed points. In this case (1) indeed reduces to transitivity. Instead of (3) we have that $Sym(A_1 \oplus A_{-1}) \to A_\bullet$ is surjective. This also implies that the action is free, for the same reasons as above.</p> Answer 3: <p>Let $V$ be an affine variety with a $\mathbb{G}_m$-action. Then you should think of $A_0$ as a good model for $V/\mathbb{G}_m$. So being connected says something like $\mathbb{G}_m$ acts transitively- at least from the point of view of functions. Being positively graded is saying something like the action of $\mathbb{G}_m$ extending to an action of the monoid scheme $\mathbb{A}^1$ (under multiplication). The last condition is more involved: The module $A_1$ is the set of $\mathbb{G}_m$-equivariant functions to $\mathbb{A}^1$ where we use the scaling action on $\mathbb{A}^1$. (So it's kind of like "linear" functions on $V$). This is a vector space over $k$, call its <em>dual</em> $W$ and view it as a variety. There's a natural equivariant map $V \to W$ given by evaluation (that's why we took the dual of the vector space of functions). We're asking that this equivariant map be an embedding. So, while $V$ is not a vector space with the standard linear action of $\mathbb{G}_m$, it can be embedded in one in some natural way.</p> Answer 4: <p>Expanding on Dylan Wilson's answer, condition (2) is indeed just a continuation of $\mathbb{G}_m$-action to an $\mathbb{A}^1$-action, since $\mathcal{O}(\mathbb{A}^1) \hookrightarrow \mathcal{O}(\mathbb{G}_m): \mathbb{k}[t] \hookrightarrow \mathbb{k}[t, t^{-1}]$. Condition (1) is trickier, since the action of $\mathbb{G}_m$ on $V$ will have inseparable orbits --- see the action on $\mathbb{A}^n$. Thus we can't think of it as transitivity (not of the group action at least, the monoid $\mathbb{A}^1$ action is transitive in a sense that any two points can be mapped to a same point). Instead we should look at it together with condition (2), which implies that the action of $\mathbb{G}_m$ has fixed points: for any $x \in V$ the point $0\cdot x \in V$ will be fixed. Here $0 \in \mathbb{A}^1$ and its action exists by (2). Condition (1) then reduces to the statement that the $\mathbb{G}_m$-action has a <em>unique</em> fixed point. On the level of functions it is given by a graded morphism $\mathcal{O}(V) \to \mathbb k$ as graded algebras. Condition (3) means that the action is <em>free</em> if we remove that single fixed point. We can see it because the action on $A_1^*$ is a free action apart from $0 \in A_1^*$ and the embedding $V \hookrightarrow A_1^*$ is $\mathbb{G}_m$-equivariant, as a morphism of graded algebras.</p> <p>The simplest example where you can see all those statements is the obvious action on $\mathbb{A}^n$. On the other hand, consider the obvious action of $\mathbb{G}_m$ on itself. It corresponds to the natural grading on $\mathbb{k}[t,t^{-1}]$. The reasons (2) fails here is obvious. Since there is no morphism $\mathbb{k}[t,t^{-1}] \to \mathbb{k}$, the action has no fixed points. In this case (1) indeed reduces to transitivity. Instead of (3) we have that $Sym(A_1 \oplus A_{-1}) \to A_\bullet$ is surjective. This also implies that the action is free, for the same reasons as above.</p>
When is a compact Lie group endowed with a left-invariant complex structure a K&#228;hler manifold of balanced manifold?
259,640
5
https://mathoverflow.net/questions/259640/when-is-a-compact-lie-group-endowed-with-a-left-invariant-complex-structure-a-k%c3%a4
<p>Let $G$ be a compact Lie group having a left-invariant complex structure $J$.</p> <p>Is there a hermitian metric $h$ in $G$, compatible with the complex structure $J$, such that $G$ is a Kähler manifold?</p> <p>In the case it is not possible to turn $G$ into a Kähler manifold, does it hold that exist a hermitian metric such that $G$ is a balanced manifold? That is, $\Delta_d = 2 \Delta_\overline{\partial}$?</p>
[ { "answer_id": 259645, "score": 7, "text": "<p>To answer the first question: Only when $G$ is abelian. In fact, if $G$ is not abelian and has dimension $2d$, then there is no element of $H^2(G,\\mathbb{R})$ whose $d$-th power is nonzero in $H^{2d}(G,\\mathbb{R})$ (just look at the representation by bi-invariant forms), and there would have to be such an element if there were going to be even a symplectic structure on $G$, much less a Kähler structure. Conversely, if $G$ is abelian, then, yes, since $G=\\mathbb{R}^{2d}/\\mathbb{Z}^{2d}$, and, for any left-invariant complex structure, $G$ is $\\mathbb{C}^d/\\Lambda$, where $\\Lambda\\subset\\mathbb{C}^d$ is a lattice (i.e., a discrete subgroup of maximal rank $2d$). Thus, there is a left-invariant Kähler structure, for example, (though there are others).</p>\n\n<p><em>Added on 21 Jan 2017:</em> As for your second question, I believe I have a partial answer, which is 'If the center of $G$ is finite (i.e., $G$ is semisimple), then $G$ does not possess a balanced metric'. Here is why: It is my understanding that, if $G$ is a compact Lie group and $J$ is a left-invariant complex structure, then $J$ is one of the complex structures constructed by Samelson and Wang. (See H.C. Wang, <em>Closed Manifolds with Homogeneous Complex Structure</em>, American Journal of Mathematics <strong>76</strong> (1954),1–32, which seems to indicate this though it doesn't state it explicitly. Also, see the article mentioned by François in the comments below.) </p>\n\n<p>In such cases, there exists a maximal toral subgroup $T\\subset G$ that is $J$-complex, and the quotient $G/T$ inherits a complex structure such that the quotient map $\\pi:G\\to G/T$ is holomorphic. Moreover, $G/T$ is not only Kähler, but an algebraic complex manifold to boot (in fact, it's the complete flag variety of $G$). </p>\n\n<p>In particular, $G/T$ contains an algebraic (complex) hypersurface $Z\\subset G/T$, whose preimage $X=\\pi^{-1}(Z)\\subset G$ is therefore a compact complex hypersurface in $G$. Now if $\\omega$ were a positive $(1,1)$-form on $G$ (with respect to the complex structure $J$) such that $\\omega^{d-1}$ were closed (where $2d$ is the dimension of $G$), then the integral of $\\omega^{d-1}$ over $X$ would be positive. In particular, it would follow that the homology class of $X$ in $H_{2d-2}(G,\\mathbb{R})$ were nonzero. </p>\n\n<p>However, when $G$ is semi-simple (i.e., when its center is finite), we know by the Whitehead Lemma that $H_2(G,\\mathbb{R}) = 0$, so Poincaré Duality implies that $H_{2d-2}(G,\\mathbb{R})=0$ as well, contradicting the fact that the real homology class of $X$ must be nonzero.</p>\n\n<p>I believe a more careful analysis of the cohomology ring of $G$ when the center of $G$ has positive dimension will lead to the result that $G$ can carry a balanced metric only when $G$ is abelian.</p>\n" }, { "answer_id": 259643, "score": 4, "text": "<p>To partial answer the first question:\nTheorem (A.Borel- R.Remmert) Let $X$ be a connected, compact, homogeneous Kaehler manifold. Then $X$ is isomorphic to the product $Y \\times A(X)$, where\n$Y=G/P$ is a flag manifold and $A(X)$ is the Albanese manifold. More precisely, $Y$ is the base of the Tits fibration and the isomorphism is given by the product of the Tits fibration map and the Albanese map.</p>\n" }, { "answer_id": 259645, "score": 7, "text": "<p>To answer the first question: Only when $G$ is abelian. In fact, if $G$ is not abelian and has dimension $2d$, then there is no element of $H^2(G,\\mathbb{R})$ whose $d$-th power is nonzero in $H^{2d}(G,\\mathbb{R})$ (just look at the representation by bi-invariant forms), and there would have to be such an element if there were going to be even a symplectic structure on $G$, much less a Kähler structure. Conversely, if $G$ is abelian, then, yes, since $G=\\mathbb{R}^{2d}/\\mathbb{Z}^{2d}$, and, for any left-invariant complex structure, $G$ is $\\mathbb{C}^d/\\Lambda$, where $\\Lambda\\subset\\mathbb{C}^d$ is a lattice (i.e., a discrete subgroup of maximal rank $2d$). Thus, there is a left-invariant Kähler structure, for example, (though there are others).</p>\n\n<p><em>Added on 21 Jan 2017:</em> As for your second question, I believe I have a partial answer, which is 'If the center of $G$ is finite (i.e., $G$ is semisimple), then $G$ does not possess a balanced metric'. Here is why: It is my understanding that, if $G$ is a compact Lie group and $J$ is a left-invariant complex structure, then $J$ is one of the complex structures constructed by Samelson and Wang. (See H.C. Wang, <em>Closed Manifolds with Homogeneous Complex Structure</em>, American Journal of Mathematics <strong>76</strong> (1954),1–32, which seems to indicate this though it doesn't state it explicitly. Also, see the article mentioned by François in the comments below.) </p>\n\n<p>In such cases, there exists a maximal toral subgroup $T\\subset G$ that is $J$-complex, and the quotient $G/T$ inherits a complex structure such that the quotient map $\\pi:G\\to G/T$ is holomorphic. Moreover, $G/T$ is not only Kähler, but an algebraic complex manifold to boot (in fact, it's the complete flag variety of $G$). </p>\n\n<p>In particular, $G/T$ contains an algebraic (complex) hypersurface $Z\\subset G/T$, whose preimage $X=\\pi^{-1}(Z)\\subset G$ is therefore a compact complex hypersurface in $G$. Now if $\\omega$ were a positive $(1,1)$-form on $G$ (with respect to the complex structure $J$) such that $\\omega^{d-1}$ were closed (where $2d$ is the dimension of $G$), then the integral of $\\omega^{d-1}$ over $X$ would be positive. In particular, it would follow that the homology class of $X$ in $H_{2d-2}(G,\\mathbb{R})$ were nonzero. </p>\n\n<p>However, when $G$ is semi-simple (i.e., when its center is finite), we know by the Whitehead Lemma that $H_2(G,\\mathbb{R}) = 0$, so Poincaré Duality implies that $H_{2d-2}(G,\\mathbb{R})=0$ as well, contradicting the fact that the real homology class of $X$ must be nonzero.</p>\n\n<p>I believe a more careful analysis of the cohomology ring of $G$ when the center of $G$ has positive dimension will lead to the result that $G$ can carry a balanced metric only when $G$ is abelian.</p>\n" }, { "answer_id": 259643, "score": 4, "text": "<p>To partial answer the first question:\nTheorem (A.Borel- R.Remmert) Let $X$ be a connected, compact, homogeneous Kaehler manifold. Then $X$ is isomorphic to the product $Y \\times A(X)$, where\n$Y=G/P$ is a flag manifold and $A(X)$ is the Albanese manifold. More precisely, $Y$ is the base of the Tits fibration and the isomorphism is given by the product of the Tits fibration map and the Albanese map.</p>\n" } ]
Question: When is a compact Lie group endowed with a left-invariant complex structure a K&#228;hler manifold of balanced manifold? Answer 1: <p>To answer the first question: Only when $G$ is abelian. In fact, if $G$ is not abelian and has dimension $2d$, then there is no element of $H^2(G,\mathbb{R})$ whose $d$-th power is nonzero in $H^{2d}(G,\mathbb{R})$ (just look at the representation by bi-invariant forms), and there would have to be such an element if there were going to be even a symplectic structure on $G$, much less a Kähler structure. Conversely, if $G$ is abelian, then, yes, since $G=\mathbb{R}^{2d}/\mathbb{Z}^{2d}$, and, for any left-invariant complex structure, $G$ is $\mathbb{C}^d/\Lambda$, where $\Lambda\subset\mathbb{C}^d$ is a lattice (i.e., a discrete subgroup of maximal rank $2d$). Thus, there is a left-invariant Kähler structure, for example, (though there are others).</p> <p><em>Added on 21 Jan 2017:</em> As for your second question, I believe I have a partial answer, which is 'If the center of $G$ is finite (i.e., $G$ is semisimple), then $G$ does not possess a balanced metric'. Here is why: It is my understanding that, if $G$ is a compact Lie group and $J$ is a left-invariant complex structure, then $J$ is one of the complex structures constructed by Samelson and Wang. (See H.C. Wang, <em>Closed Manifolds with Homogeneous Complex Structure</em>, American Journal of Mathematics <strong>76</strong> (1954),1–32, which seems to indicate this though it doesn't state it explicitly. Also, see the article mentioned by François in the comments below.) </p> <p>In such cases, there exists a maximal toral subgroup $T\subset G$ that is $J$-complex, and the quotient $G/T$ inherits a complex structure such that the quotient map $\pi:G\to G/T$ is holomorphic. Moreover, $G/T$ is not only Kähler, but an algebraic complex manifold to boot (in fact, it's the complete flag variety of $G$). </p> <p>In particular, $G/T$ contains an algebraic (complex) hypersurface $Z\subset G/T$, whose preimage $X=\pi^{-1}(Z)\subset G$ is therefore a compact complex hypersurface in $G$. Now if $\omega$ were a positive $(1,1)$-form on $G$ (with respect to the complex structure $J$) such that $\omega^{d-1}$ were closed (where $2d$ is the dimension of $G$), then the integral of $\omega^{d-1}$ over $X$ would be positive. In particular, it would follow that the homology class of $X$ in $H_{2d-2}(G,\mathbb{R})$ were nonzero. </p> <p>However, when $G$ is semi-simple (i.e., when its center is finite), we know by the Whitehead Lemma that $H_2(G,\mathbb{R}) = 0$, so Poincaré Duality implies that $H_{2d-2}(G,\mathbb{R})=0$ as well, contradicting the fact that the real homology class of $X$ must be nonzero.</p> <p>I believe a more careful analysis of the cohomology ring of $G$ when the center of $G$ has positive dimension will lead to the result that $G$ can carry a balanced metric only when $G$ is abelian.</p> Answer 2: <p>To partial answer the first question: Theorem (A.Borel- R.Remmert) Let $X$ be a connected, compact, homogeneous Kaehler manifold. Then $X$ is isomorphic to the product $Y \times A(X)$, where $Y=G/P$ is a flag manifold and $A(X)$ is the Albanese manifold. More precisely, $Y$ is the base of the Tits fibration and the isomorphism is given by the product of the Tits fibration map and the Albanese map.</p> Answer 3: <p>To answer the first question: Only when $G$ is abelian. In fact, if $G$ is not abelian and has dimension $2d$, then there is no element of $H^2(G,\mathbb{R})$ whose $d$-th power is nonzero in $H^{2d}(G,\mathbb{R})$ (just look at the representation by bi-invariant forms), and there would have to be such an element if there were going to be even a symplectic structure on $G$, much less a Kähler structure. Conversely, if $G$ is abelian, then, yes, since $G=\mathbb{R}^{2d}/\mathbb{Z}^{2d}$, and, for any left-invariant complex structure, $G$ is $\mathbb{C}^d/\Lambda$, where $\Lambda\subset\mathbb{C}^d$ is a lattice (i.e., a discrete subgroup of maximal rank $2d$). Thus, there is a left-invariant Kähler structure, for example, (though there are others).</p> <p><em>Added on 21 Jan 2017:</em> As for your second question, I believe I have a partial answer, which is 'If the center of $G$ is finite (i.e., $G$ is semisimple), then $G$ does not possess a balanced metric'. Here is why: It is my understanding that, if $G$ is a compact Lie group and $J$ is a left-invariant complex structure, then $J$ is one of the complex structures constructed by Samelson and Wang. (See H.C. Wang, <em>Closed Manifolds with Homogeneous Complex Structure</em>, American Journal of Mathematics <strong>76</strong> (1954),1–32, which seems to indicate this though it doesn't state it explicitly. Also, see the article mentioned by François in the comments below.) </p> <p>In such cases, there exists a maximal toral subgroup $T\subset G$ that is $J$-complex, and the quotient $G/T$ inherits a complex structure such that the quotient map $\pi:G\to G/T$ is holomorphic. Moreover, $G/T$ is not only Kähler, but an algebraic complex manifold to boot (in fact, it's the complete flag variety of $G$). </p> <p>In particular, $G/T$ contains an algebraic (complex) hypersurface $Z\subset G/T$, whose preimage $X=\pi^{-1}(Z)\subset G$ is therefore a compact complex hypersurface in $G$. Now if $\omega$ were a positive $(1,1)$-form on $G$ (with respect to the complex structure $J$) such that $\omega^{d-1}$ were closed (where $2d$ is the dimension of $G$), then the integral of $\omega^{d-1}$ over $X$ would be positive. In particular, it would follow that the homology class of $X$ in $H_{2d-2}(G,\mathbb{R})$ were nonzero. </p> <p>However, when $G$ is semi-simple (i.e., when its center is finite), we know by the Whitehead Lemma that $H_2(G,\mathbb{R}) = 0$, so Poincaré Duality implies that $H_{2d-2}(G,\mathbb{R})=0$ as well, contradicting the fact that the real homology class of $X$ must be nonzero.</p> <p>I believe a more careful analysis of the cohomology ring of $G$ when the center of $G$ has positive dimension will lead to the result that $G$ can carry a balanced metric only when $G$ is abelian.</p> Answer 4: <p>To partial answer the first question: Theorem (A.Borel- R.Remmert) Let $X$ be a connected, compact, homogeneous Kaehler manifold. Then $X$ is isomorphic to the product $Y \times A(X)$, where $Y=G/P$ is a flag manifold and $A(X)$ is the Albanese manifold. More precisely, $Y$ is the base of the Tits fibration and the isomorphism is given by the product of the Tits fibration map and the Albanese map.</p>
Proper coverings as a quotient of the coverings
260,303
5
https://mathoverflow.net/questions/260303/proper-coverings-as-a-quotient-of-the-coverings
<p>Let $X\neq \emptyset$ be a set. We say that $U\subseteq {\cal P}(X)\setminus \{\emptyset\}$ is a <em>covering</em> of $X$ if $\bigcup U = X$. We call a covering <em>proper</em> if for $a\neq b\in U$ we have $a\not\subseteq b$.</p> <p>Let $\text{Cov}(X)$ denote the collection of all coverings of $X$. For $A, B\in \text{Cov}(X)$ we set $A\leq B$ if $A$ <em>refines</em> $B$, that is for all $a\in A$ there is $b\in B$ such that $a\subseteq b$.</p> <p>This relation defines a pre-order on $\text{Cov}(X)$; it is easily seen that it is not anti-symmetric. However, the refinement relation is anti-symmetric on $\text{PropCov}(X)$, the set of proper coverings.</p> <p>On $\text{Cov}(X)$ we set $A \simeq B$ if $A\leq B$ and $B\leq A$. It is easy to verify that the relation $\leq_q $ on $\text{Cov}(X)/\simeq$ defined by $$[A]_\simeq \leq_q [B]_\simeq \text{ if and only if } A \leq B \text{ in Cov}(X)$$ is a partially ordered set.</p> <p>Are $(\text{Cov}(X)/\simeq, \leq_q)$ and $(\text{PropCov}(X),\leq)$ isomorphic as posets?</p>
[ { "answer_id": 312980, "score": 3, "text": "<p><strong>Are <span class=\"math-container\">$(\\text{Cov}(X)/\\simeq, \\leq_q)$</span> and\n<span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span> isomorphic as posets?</strong>\n<p></p>\n\n<p>If <span class=\"math-container\">$X$</span> is finite, then <span class=\"math-container\">$(\\text{Cov}(X)/\\simeq, \\leq_q)$</span>\nand <span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span> \nare isomorphic, but if <span class=\"math-container\">$X$</span> is infinite they are not.<p></p>\n\n<p>To see why they are isomorphic when <span class=\"math-container\">$X$</span> is finite it is enough to\nnote that, in the case of finite <span class=\"math-container\">$X$</span>, each <span class=\"math-container\">$\\simeq$</span>-class\nof covers has the property that all members share\nthe same maximal elements. Moreover, this common set \nof maximal elements constitutes a proper cover,\nand it is the unique proper cover in this <span class=\"math-container\">$\\simeq$</span>-class.\nThis bijection between <span class=\"math-container\">$\\simeq$</span>-classes of covers\nand proper covers preserves and reflects order, hence establishes\nan isomorphism between <span class=\"math-container\">$(\\text{Cov}(X)/\\simeq, \\leq_q)$</span> and\n<span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span>. <p></p>\n\n<p>To see why <span class=\"math-container\">$(\\text{Cov}(X)/\\simeq, \\leq_q)$</span>\nand <span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span> \nare not isomorphic whene <span class=\"math-container\">$X$</span> is infinite,\nit will suffice to point out that \n<span class=\"math-container\">$(\\text{Cov}(X)/\\simeq, \\leq_q)$</span> is a lattice\nwhile <span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span> is not.\n<p></p>\n\n<p>I first argue the easy part: \n<span class=\"math-container\">$(\\text{Cov}(X)/\\simeq, \\leq_q)$</span>\nis a lattice.</p>\n\n<p>Consider the following closure operator\non the set of covers:\ngiven <span class=\"math-container\">$B\\in \\text{Cov}(X)$</span>, let <span class=\"math-container\">$B^* = \\bigcup_{A\\leq B} A$</span>.\nIt is easy to see that <span class=\"math-container\">$B^*$</span> is a cover of <span class=\"math-container\">$X$</span>, and that\nan alternative definition for it is\n<span class=\"math-container\">$$\nB^* = (\\bigcup_{b\\in B} {\\mathcal P}(b))\\setminus \\{\\emptyset\\}.\n$$</span>\nMoreover, <span class=\"math-container\">$A\\leq B$</span> holds iff <span class=\"math-container\">$A\\subseteq B^*$</span> iff <span class=\"math-container\">$A^*\\subseteq B^*$</span>,\nso <span class=\"math-container\">$A\\simeq B$</span> iff <span class=\"math-container\">$A^*=B^*$</span>.</p>\n\n<p>Now, a subset <span class=\"math-container\">$Z\\subseteq {\\mathcal P}(X)$</span> is of the form <span class=\"math-container\">$Z=B^*$</span>\niff (i) <span class=\"math-container\">$\\emptyset\\notin Z$</span> and (ii) <span class=\"math-container\">$Z\\cup\\{\\emptyset\\}$</span>\nis an order ideal of <span class=\"math-container\">${\\mathcal P}(X)$</span> which contains\nall the singleton subsets of <span class=\"math-container\">$X$</span>. This is enough to show that \n<span class=\"math-container\">$(\\text{Cov}(X)/\\simeq, \\leq_q)$</span> is isomorphic to the\ninterval of the lattice of order ideals of <span class=\"math-container\">${\\mathcal P}(X)$</span>\nabove the order ideal consisting of singleton sets.\nThis is a distributive, algebraic lattice.</p>\n\n<p><p></p>\n\n<p>Now the less-easy part. I argue that \n<span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span> is not a lattice.\nIn fact, I will write out only the case\n<span class=\"math-container\">$X=\\omega$</span>, but afterwards I will explain how to extend the argument to\nany superset of <span class=\"math-container\">$\\omega$</span>.\n<p></p>\n\n<p>Let's name some subsets of <span class=\"math-container\">$\\omega$</span>:<br>\n<span class=\"math-container\">$E$</span>: the set <span class=\"math-container\">$\\{0,2,4,\\ldots\\}$</span> \nof all even natural numbers.<br>\n<span class=\"math-container\">$E_n$</span>: the set <span class=\"math-container\">$\\{0,2,4,\\ldots,2n\\}$</span> \nof even numbers up to <span class=\"math-container\">$2n$</span>.<br>\n<span class=\"math-container\">$E_n^*$</span>: the set <span class=\"math-container\">$\\{0,2,4,\\ldots,2n,2n+1\\}$</span> \nof even numbers up to <span class=\"math-container\">$2n$</span>\nalong with one odd number <span class=\"math-container\">$2n+1$</span>.<p></p>\n\n<p>Now let's name some proper covers of <span class=\"math-container\">$\\omega$</span>:<br>\n<span class=\"math-container\">$C_E$</span>: the proper cover consisting of <span class=\"math-container\">$E$</span> and all odd singletons.<br>\n<span class=\"math-container\">$C_{E^*}$</span>: The proper cover consisting of all <span class=\"math-container\">$E_n^*$</span>'s.<br>\n<span class=\"math-container\">$C_n$</span>: The proper cover including <span class=\"math-container\">$E_n$</span> and all singletons\nof elements from <span class=\"math-container\">$\\omega \\setminus E_n$</span>.<p></p>\n\n<p>It is not hard to see that<br>\n(1) <span class=\"math-container\">$C_E$</span> and <span class=\"math-container\">$C_{E^*}$</span> are incomparable under <span class=\"math-container\">$\\leq$</span>.\n<br>\n(2) <span class=\"math-container\">$C_n\\leq C_E$</span> and <span class=\"math-container\">$C_n\\leq C_{E^*}$</span> for all <span class=\"math-container\">$n$</span>.<br></p>\n\n<p>My goal is to show that <span class=\"math-container\">$C_E$</span> and <span class=\"math-container\">$C_{E^*}$</span> do not have a\nmeet (= greatest lower bound) in <span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span>.<p></p>\n\n<p>Suppose that <span class=\"math-container\">$D$</span> is a proper cover of <span class=\"math-container\">$\\omega$</span>\nthat is a candidate for the meet \nof <span class=\"math-container\">$C_E$</span> and <span class=\"math-container\">$C_{E^*}$</span> in <span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span>.\nThen <span class=\"math-container\">$D\\leq C_E$</span>, <span class=\"math-container\">$D\\leq C_{E^*}$</span>, and <span class=\"math-container\">$D$</span> is above all other\ncommon lower bounds of\n<span class=\"math-container\">$C_E$</span> and <span class=\"math-container\">$C_{E^*}$</span>, for example <span class=\"math-container\">$C_n\\leq D$</span> for all <span class=\"math-container\">$n$</span>.</p>\n\n<p>Since <span class=\"math-container\">$E_n=\\{0,2,4,\\ldots,2n\\}\\in C_n$</span>,\nthere must be a set <span class=\"math-container\">$d_n\\in D$</span> such that <span class=\"math-container\">$E_n\\subseteq d_n$</span>.\nSince <span class=\"math-container\">$D\\leq C_E$</span>, it follows that <span class=\"math-container\">$d_n$</span> consists of even\nintegers.\nSince <span class=\"math-container\">$D\\leq C_{E^*}$</span>, it follows that <span class=\"math-container\">$d_n$</span> is a finite set.\nThus <span class=\"math-container\">$d_n$</span> is properly contained in some\n<span class=\"math-container\">$E_m=\\{0,2,\\ldots,2m\\}$</span> for sufficiently\nlarge <span class=\"math-container\">$m$</span>. But now repeat this argument to find\nsome <span class=\"math-container\">$d_m\\in D$</span> containing <span class=\"math-container\">$E_m$</span>. We now have\n<span class=\"math-container\">$d_n\\subsetneq E_m \\subseteq d_m$</span>, contradicting the\nproperness of <span class=\"math-container\">$D$</span>.\n\\\\\\\\\\</p>\n\n<p>Finally, to extend the\nproof that <span class=\"math-container\">$(\\text{PropCov}(\\omega),\\leq)$</span> is not a lattice\nto a proof that <span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span> is not a lattice,\nfor any superset <span class=\"math-container\">$X\\supseteq \\omega$</span>, repeat the argument\nabove, but extend all the covers of <span class=\"math-container\">$\\omega$</span> which were used\nin the argument to covers\nof <span class=\"math-container\">$X$</span> by adding to those covers\nall singletons <span class=\"math-container\">$\\{x\\}$</span>, <span class=\"math-container\">$x\\in X\\setminus\\omega$</span>.</p>\n" }, { "answer_id": 260315, "score": 2, "text": "<p>I don't think they are isomorphic. Let $X=\\mathbb N$ and let $A$ be a covering consisting of the sets of the form\n$$ a_n = \\{ 0, \\dots, n \\} $$\nfor every $n\\in \\mathbb N$.</p>\n\n<p>Then, $A$ is not $\\simeq$-equivalent to any proper covering.</p>\n" }, { "answer_id": 312980, "score": 3, "text": "<p><strong>Are <span class=\"math-container\">$(\\text{Cov}(X)/\\simeq, \\leq_q)$</span> and\n<span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span> isomorphic as posets?</strong>\n<p></p>\n\n<p>If <span class=\"math-container\">$X$</span> is finite, then <span class=\"math-container\">$(\\text{Cov}(X)/\\simeq, \\leq_q)$</span>\nand <span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span> \nare isomorphic, but if <span class=\"math-container\">$X$</span> is infinite they are not.<p></p>\n\n<p>To see why they are isomorphic when <span class=\"math-container\">$X$</span> is finite it is enough to\nnote that, in the case of finite <span class=\"math-container\">$X$</span>, each <span class=\"math-container\">$\\simeq$</span>-class\nof covers has the property that all members share\nthe same maximal elements. Moreover, this common set \nof maximal elements constitutes a proper cover,\nand it is the unique proper cover in this <span class=\"math-container\">$\\simeq$</span>-class.\nThis bijection between <span class=\"math-container\">$\\simeq$</span>-classes of covers\nand proper covers preserves and reflects order, hence establishes\nan isomorphism between <span class=\"math-container\">$(\\text{Cov}(X)/\\simeq, \\leq_q)$</span> and\n<span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span>. <p></p>\n\n<p>To see why <span class=\"math-container\">$(\\text{Cov}(X)/\\simeq, \\leq_q)$</span>\nand <span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span> \nare not isomorphic whene <span class=\"math-container\">$X$</span> is infinite,\nit will suffice to point out that \n<span class=\"math-container\">$(\\text{Cov}(X)/\\simeq, \\leq_q)$</span> is a lattice\nwhile <span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span> is not.\n<p></p>\n\n<p>I first argue the easy part: \n<span class=\"math-container\">$(\\text{Cov}(X)/\\simeq, \\leq_q)$</span>\nis a lattice.</p>\n\n<p>Consider the following closure operator\non the set of covers:\ngiven <span class=\"math-container\">$B\\in \\text{Cov}(X)$</span>, let <span class=\"math-container\">$B^* = \\bigcup_{A\\leq B} A$</span>.\nIt is easy to see that <span class=\"math-container\">$B^*$</span> is a cover of <span class=\"math-container\">$X$</span>, and that\nan alternative definition for it is\n<span class=\"math-container\">$$\nB^* = (\\bigcup_{b\\in B} {\\mathcal P}(b))\\setminus \\{\\emptyset\\}.\n$$</span>\nMoreover, <span class=\"math-container\">$A\\leq B$</span> holds iff <span class=\"math-container\">$A\\subseteq B^*$</span> iff <span class=\"math-container\">$A^*\\subseteq B^*$</span>,\nso <span class=\"math-container\">$A\\simeq B$</span> iff <span class=\"math-container\">$A^*=B^*$</span>.</p>\n\n<p>Now, a subset <span class=\"math-container\">$Z\\subseteq {\\mathcal P}(X)$</span> is of the form <span class=\"math-container\">$Z=B^*$</span>\niff (i) <span class=\"math-container\">$\\emptyset\\notin Z$</span> and (ii) <span class=\"math-container\">$Z\\cup\\{\\emptyset\\}$</span>\nis an order ideal of <span class=\"math-container\">${\\mathcal P}(X)$</span> which contains\nall the singleton subsets of <span class=\"math-container\">$X$</span>. This is enough to show that \n<span class=\"math-container\">$(\\text{Cov}(X)/\\simeq, \\leq_q)$</span> is isomorphic to the\ninterval of the lattice of order ideals of <span class=\"math-container\">${\\mathcal P}(X)$</span>\nabove the order ideal consisting of singleton sets.\nThis is a distributive, algebraic lattice.</p>\n\n<p><p></p>\n\n<p>Now the less-easy part. I argue that \n<span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span> is not a lattice.\nIn fact, I will write out only the case\n<span class=\"math-container\">$X=\\omega$</span>, but afterwards I will explain how to extend the argument to\nany superset of <span class=\"math-container\">$\\omega$</span>.\n<p></p>\n\n<p>Let's name some subsets of <span class=\"math-container\">$\\omega$</span>:<br>\n<span class=\"math-container\">$E$</span>: the set <span class=\"math-container\">$\\{0,2,4,\\ldots\\}$</span> \nof all even natural numbers.<br>\n<span class=\"math-container\">$E_n$</span>: the set <span class=\"math-container\">$\\{0,2,4,\\ldots,2n\\}$</span> \nof even numbers up to <span class=\"math-container\">$2n$</span>.<br>\n<span class=\"math-container\">$E_n^*$</span>: the set <span class=\"math-container\">$\\{0,2,4,\\ldots,2n,2n+1\\}$</span> \nof even numbers up to <span class=\"math-container\">$2n$</span>\nalong with one odd number <span class=\"math-container\">$2n+1$</span>.<p></p>\n\n<p>Now let's name some proper covers of <span class=\"math-container\">$\\omega$</span>:<br>\n<span class=\"math-container\">$C_E$</span>: the proper cover consisting of <span class=\"math-container\">$E$</span> and all odd singletons.<br>\n<span class=\"math-container\">$C_{E^*}$</span>: The proper cover consisting of all <span class=\"math-container\">$E_n^*$</span>'s.<br>\n<span class=\"math-container\">$C_n$</span>: The proper cover including <span class=\"math-container\">$E_n$</span> and all singletons\nof elements from <span class=\"math-container\">$\\omega \\setminus E_n$</span>.<p></p>\n\n<p>It is not hard to see that<br>\n(1) <span class=\"math-container\">$C_E$</span> and <span class=\"math-container\">$C_{E^*}$</span> are incomparable under <span class=\"math-container\">$\\leq$</span>.\n<br>\n(2) <span class=\"math-container\">$C_n\\leq C_E$</span> and <span class=\"math-container\">$C_n\\leq C_{E^*}$</span> for all <span class=\"math-container\">$n$</span>.<br></p>\n\n<p>My goal is to show that <span class=\"math-container\">$C_E$</span> and <span class=\"math-container\">$C_{E^*}$</span> do not have a\nmeet (= greatest lower bound) in <span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span>.<p></p>\n\n<p>Suppose that <span class=\"math-container\">$D$</span> is a proper cover of <span class=\"math-container\">$\\omega$</span>\nthat is a candidate for the meet \nof <span class=\"math-container\">$C_E$</span> and <span class=\"math-container\">$C_{E^*}$</span> in <span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span>.\nThen <span class=\"math-container\">$D\\leq C_E$</span>, <span class=\"math-container\">$D\\leq C_{E^*}$</span>, and <span class=\"math-container\">$D$</span> is above all other\ncommon lower bounds of\n<span class=\"math-container\">$C_E$</span> and <span class=\"math-container\">$C_{E^*}$</span>, for example <span class=\"math-container\">$C_n\\leq D$</span> for all <span class=\"math-container\">$n$</span>.</p>\n\n<p>Since <span class=\"math-container\">$E_n=\\{0,2,4,\\ldots,2n\\}\\in C_n$</span>,\nthere must be a set <span class=\"math-container\">$d_n\\in D$</span> such that <span class=\"math-container\">$E_n\\subseteq d_n$</span>.\nSince <span class=\"math-container\">$D\\leq C_E$</span>, it follows that <span class=\"math-container\">$d_n$</span> consists of even\nintegers.\nSince <span class=\"math-container\">$D\\leq C_{E^*}$</span>, it follows that <span class=\"math-container\">$d_n$</span> is a finite set.\nThus <span class=\"math-container\">$d_n$</span> is properly contained in some\n<span class=\"math-container\">$E_m=\\{0,2,\\ldots,2m\\}$</span> for sufficiently\nlarge <span class=\"math-container\">$m$</span>. But now repeat this argument to find\nsome <span class=\"math-container\">$d_m\\in D$</span> containing <span class=\"math-container\">$E_m$</span>. We now have\n<span class=\"math-container\">$d_n\\subsetneq E_m \\subseteq d_m$</span>, contradicting the\nproperness of <span class=\"math-container\">$D$</span>.\n\\\\\\\\\\</p>\n\n<p>Finally, to extend the\nproof that <span class=\"math-container\">$(\\text{PropCov}(\\omega),\\leq)$</span> is not a lattice\nto a proof that <span class=\"math-container\">$(\\text{PropCov}(X),\\leq)$</span> is not a lattice,\nfor any superset <span class=\"math-container\">$X\\supseteq \\omega$</span>, repeat the argument\nabove, but extend all the covers of <span class=\"math-container\">$\\omega$</span> which were used\nin the argument to covers\nof <span class=\"math-container\">$X$</span> by adding to those covers\nall singletons <span class=\"math-container\">$\\{x\\}$</span>, <span class=\"math-container\">$x\\in X\\setminus\\omega$</span>.</p>\n" }, { "answer_id": 260315, "score": 2, "text": "<p>I don't think they are isomorphic. Let $X=\\mathbb N$ and let $A$ be a covering consisting of the sets of the form\n$$ a_n = \\{ 0, \\dots, n \\} $$\nfor every $n\\in \\mathbb N$.</p>\n\n<p>Then, $A$ is not $\\simeq$-equivalent to any proper covering.</p>\n" } ]
Question: Proper coverings as a quotient of the coverings Answer 1: <p><strong>Are <span class="math-container">$(\text{Cov}(X)/\simeq, \leq_q)$</span> and <span class="math-container">$(\text{PropCov}(X),\leq)$</span> isomorphic as posets?</strong> <p></p> <p>If <span class="math-container">$X$</span> is finite, then <span class="math-container">$(\text{Cov}(X)/\simeq, \leq_q)$</span> and <span class="math-container">$(\text{PropCov}(X),\leq)$</span> are isomorphic, but if <span class="math-container">$X$</span> is infinite they are not.<p></p> <p>To see why they are isomorphic when <span class="math-container">$X$</span> is finite it is enough to note that, in the case of finite <span class="math-container">$X$</span>, each <span class="math-container">$\simeq$</span>-class of covers has the property that all members share the same maximal elements. Moreover, this common set of maximal elements constitutes a proper cover, and it is the unique proper cover in this <span class="math-container">$\simeq$</span>-class. This bijection between <span class="math-container">$\simeq$</span>-classes of covers and proper covers preserves and reflects order, hence establishes an isomorphism between <span class="math-container">$(\text{Cov}(X)/\simeq, \leq_q)$</span> and <span class="math-container">$(\text{PropCov}(X),\leq)$</span>. <p></p> <p>To see why <span class="math-container">$(\text{Cov}(X)/\simeq, \leq_q)$</span> and <span class="math-container">$(\text{PropCov}(X),\leq)$</span> are not isomorphic whene <span class="math-container">$X$</span> is infinite, it will suffice to point out that <span class="math-container">$(\text{Cov}(X)/\simeq, \leq_q)$</span> is a lattice while <span class="math-container">$(\text{PropCov}(X),\leq)$</span> is not. <p></p> <p>I first argue the easy part: <span class="math-container">$(\text{Cov}(X)/\simeq, \leq_q)$</span> is a lattice.</p> <p>Consider the following closure operator on the set of covers: given <span class="math-container">$B\in \text{Cov}(X)$</span>, let <span class="math-container">$B^* = \bigcup_{A\leq B} A$</span>. It is easy to see that <span class="math-container">$B^*$</span> is a cover of <span class="math-container">$X$</span>, and that an alternative definition for it is <span class="math-container">$$ B^* = (\bigcup_{b\in B} {\mathcal P}(b))\setminus \{\emptyset\}. $$</span> Moreover, <span class="math-container">$A\leq B$</span> holds iff <span class="math-container">$A\subseteq B^*$</span> iff <span class="math-container">$A^*\subseteq B^*$</span>, so <span class="math-container">$A\simeq B$</span> iff <span class="math-container">$A^*=B^*$</span>.</p> <p>Now, a subset <span class="math-container">$Z\subseteq {\mathcal P}(X)$</span> is of the form <span class="math-container">$Z=B^*$</span> iff (i) <span class="math-container">$\emptyset\notin Z$</span> and (ii) <span class="math-container">$Z\cup\{\emptyset\}$</span> is an order ideal of <span class="math-container">${\mathcal P}(X)$</span> which contains all the singleton subsets of <span class="math-container">$X$</span>. This is enough to show that <span class="math-container">$(\text{Cov}(X)/\simeq, \leq_q)$</span> is isomorphic to the interval of the lattice of order ideals of <span class="math-container">${\mathcal P}(X)$</span> above the order ideal consisting of singleton sets. This is a distributive, algebraic lattice.</p> <p><p></p> <p>Now the less-easy part. I argue that <span class="math-container">$(\text{PropCov}(X),\leq)$</span> is not a lattice. In fact, I will write out only the case <span class="math-container">$X=\omega$</span>, but afterwards I will explain how to extend the argument to any superset of <span class="math-container">$\omega$</span>. <p></p> <p>Let's name some subsets of <span class="math-container">$\omega$</span>:<br> <span class="math-container">$E$</span>: the set <span class="math-container">$\{0,2,4,\ldots\}$</span> of all even natural numbers.<br> <span class="math-container">$E_n$</span>: the set <span class="math-container">$\{0,2,4,\ldots,2n\}$</span> of even numbers up to <span class="math-container">$2n$</span>.<br> <span class="math-container">$E_n^*$</span>: the set <span class="math-container">$\{0,2,4,\ldots,2n,2n+1\}$</span> of even numbers up to <span class="math-container">$2n$</span> along with one odd number <span class="math-container">$2n+1$</span>.<p></p> <p>Now let's name some proper covers of <span class="math-container">$\omega$</span>:<br> <span class="math-container">$C_E$</span>: the proper cover consisting of <span class="math-container">$E$</span> and all odd singletons.<br> <span class="math-container">$C_{E^*}$</span>: The proper cover consisting of all <span class="math-container">$E_n^*$</span>'s.<br> <span class="math-container">$C_n$</span>: The proper cover including <span class="math-container">$E_n$</span> and all singletons of elements from <span class="math-container">$\omega \setminus E_n$</span>.<p></p> <p>It is not hard to see that<br> (1) <span class="math-container">$C_E$</span> and <span class="math-container">$C_{E^*}$</span> are incomparable under <span class="math-container">$\leq$</span>. <br> (2) <span class="math-container">$C_n\leq C_E$</span> and <span class="math-container">$C_n\leq C_{E^*}$</span> for all <span class="math-container">$n$</span>.<br></p> <p>My goal is to show that <span class="math-container">$C_E$</span> and <span class="math-container">$C_{E^*}$</span> do not have a meet (= greatest lower bound) in <span class="math-container">$(\text{PropCov}(X),\leq)$</span>.<p></p> <p>Suppose that <span class="math-container">$D$</span> is a proper cover of <span class="math-container">$\omega$</span> that is a candidate for the meet of <span class="math-container">$C_E$</span> and <span class="math-container">$C_{E^*}$</span> in <span class="math-container">$(\text{PropCov}(X),\leq)$</span>. Then <span class="math-container">$D\leq C_E$</span>, <span class="math-container">$D\leq C_{E^*}$</span>, and <span class="math-container">$D$</span> is above all other common lower bounds of <span class="math-container">$C_E$</span> and <span class="math-container">$C_{E^*}$</span>, for example <span class="math-container">$C_n\leq D$</span> for all <span class="math-container">$n$</span>.</p> <p>Since <span class="math-container">$E_n=\{0,2,4,\ldots,2n\}\in C_n$</span>, there must be a set <span class="math-container">$d_n\in D$</span> such that <span class="math-container">$E_n\subseteq d_n$</span>. Since <span class="math-container">$D\leq C_E$</span>, it follows that <span class="math-container">$d_n$</span> consists of even integers. Since <span class="math-container">$D\leq C_{E^*}$</span>, it follows that <span class="math-container">$d_n$</span> is a finite set. Thus <span class="math-container">$d_n$</span> is properly contained in some <span class="math-container">$E_m=\{0,2,\ldots,2m\}$</span> for sufficiently large <span class="math-container">$m$</span>. But now repeat this argument to find some <span class="math-container">$d_m\in D$</span> containing <span class="math-container">$E_m$</span>. We now have <span class="math-container">$d_n\subsetneq E_m \subseteq d_m$</span>, contradicting the properness of <span class="math-container">$D$</span>. \\\\\</p> <p>Finally, to extend the proof that <span class="math-container">$(\text{PropCov}(\omega),\leq)$</span> is not a lattice to a proof that <span class="math-container">$(\text{PropCov}(X),\leq)$</span> is not a lattice, for any superset <span class="math-container">$X\supseteq \omega$</span>, repeat the argument above, but extend all the covers of <span class="math-container">$\omega$</span> which were used in the argument to covers of <span class="math-container">$X$</span> by adding to those covers all singletons <span class="math-container">$\{x\}$</span>, <span class="math-container">$x\in X\setminus\omega$</span>.</p> Answer 2: <p>I don't think they are isomorphic. Let $X=\mathbb N$ and let $A$ be a covering consisting of the sets of the form $$ a_n = \{ 0, \dots, n \} $$ for every $n\in \mathbb N$.</p> <p>Then, $A$ is not $\simeq$-equivalent to any proper covering.</p> Answer 3: <p><strong>Are <span class="math-container">$(\text{Cov}(X)/\simeq, \leq_q)$</span> and <span class="math-container">$(\text{PropCov}(X),\leq)$</span> isomorphic as posets?</strong> <p></p> <p>If <span class="math-container">$X$</span> is finite, then <span class="math-container">$(\text{Cov}(X)/\simeq, \leq_q)$</span> and <span class="math-container">$(\text{PropCov}(X),\leq)$</span> are isomorphic, but if <span class="math-container">$X$</span> is infinite they are not.<p></p> <p>To see why they are isomorphic when <span class="math-container">$X$</span> is finite it is enough to note that, in the case of finite <span class="math-container">$X$</span>, each <span class="math-container">$\simeq$</span>-class of covers has the property that all members share the same maximal elements. Moreover, this common set of maximal elements constitutes a proper cover, and it is the unique proper cover in this <span class="math-container">$\simeq$</span>-class. This bijection between <span class="math-container">$\simeq$</span>-classes of covers and proper covers preserves and reflects order, hence establishes an isomorphism between <span class="math-container">$(\text{Cov}(X)/\simeq, \leq_q)$</span> and <span class="math-container">$(\text{PropCov}(X),\leq)$</span>. <p></p> <p>To see why <span class="math-container">$(\text{Cov}(X)/\simeq, \leq_q)$</span> and <span class="math-container">$(\text{PropCov}(X),\leq)$</span> are not isomorphic whene <span class="math-container">$X$</span> is infinite, it will suffice to point out that <span class="math-container">$(\text{Cov}(X)/\simeq, \leq_q)$</span> is a lattice while <span class="math-container">$(\text{PropCov}(X),\leq)$</span> is not. <p></p> <p>I first argue the easy part: <span class="math-container">$(\text{Cov}(X)/\simeq, \leq_q)$</span> is a lattice.</p> <p>Consider the following closure operator on the set of covers: given <span class="math-container">$B\in \text{Cov}(X)$</span>, let <span class="math-container">$B^* = \bigcup_{A\leq B} A$</span>. It is easy to see that <span class="math-container">$B^*$</span> is a cover of <span class="math-container">$X$</span>, and that an alternative definition for it is <span class="math-container">$$ B^* = (\bigcup_{b\in B} {\mathcal P}(b))\setminus \{\emptyset\}. $$</span> Moreover, <span class="math-container">$A\leq B$</span> holds iff <span class="math-container">$A\subseteq B^*$</span> iff <span class="math-container">$A^*\subseteq B^*$</span>, so <span class="math-container">$A\simeq B$</span> iff <span class="math-container">$A^*=B^*$</span>.</p> <p>Now, a subset <span class="math-container">$Z\subseteq {\mathcal P}(X)$</span> is of the form <span class="math-container">$Z=B^*$</span> iff (i) <span class="math-container">$\emptyset\notin Z$</span> and (ii) <span class="math-container">$Z\cup\{\emptyset\}$</span> is an order ideal of <span class="math-container">${\mathcal P}(X)$</span> which contains all the singleton subsets of <span class="math-container">$X$</span>. This is enough to show that <span class="math-container">$(\text{Cov}(X)/\simeq, \leq_q)$</span> is isomorphic to the interval of the lattice of order ideals of <span class="math-container">${\mathcal P}(X)$</span> above the order ideal consisting of singleton sets. This is a distributive, algebraic lattice.</p> <p><p></p> <p>Now the less-easy part. I argue that <span class="math-container">$(\text{PropCov}(X),\leq)$</span> is not a lattice. In fact, I will write out only the case <span class="math-container">$X=\omega$</span>, but afterwards I will explain how to extend the argument to any superset of <span class="math-container">$\omega$</span>. <p></p> <p>Let's name some subsets of <span class="math-container">$\omega$</span>:<br> <span class="math-container">$E$</span>: the set <span class="math-container">$\{0,2,4,\ldots\}$</span> of all even natural numbers.<br> <span class="math-container">$E_n$</span>: the set <span class="math-container">$\{0,2,4,\ldots,2n\}$</span> of even numbers up to <span class="math-container">$2n$</span>.<br> <span class="math-container">$E_n^*$</span>: the set <span class="math-container">$\{0,2,4,\ldots,2n,2n+1\}$</span> of even numbers up to <span class="math-container">$2n$</span> along with one odd number <span class="math-container">$2n+1$</span>.<p></p> <p>Now let's name some proper covers of <span class="math-container">$\omega$</span>:<br> <span class="math-container">$C_E$</span>: the proper cover consisting of <span class="math-container">$E$</span> and all odd singletons.<br> <span class="math-container">$C_{E^*}$</span>: The proper cover consisting of all <span class="math-container">$E_n^*$</span>'s.<br> <span class="math-container">$C_n$</span>: The proper cover including <span class="math-container">$E_n$</span> and all singletons of elements from <span class="math-container">$\omega \setminus E_n$</span>.<p></p> <p>It is not hard to see that<br> (1) <span class="math-container">$C_E$</span> and <span class="math-container">$C_{E^*}$</span> are incomparable under <span class="math-container">$\leq$</span>. <br> (2) <span class="math-container">$C_n\leq C_E$</span> and <span class="math-container">$C_n\leq C_{E^*}$</span> for all <span class="math-container">$n$</span>.<br></p> <p>My goal is to show that <span class="math-container">$C_E$</span> and <span class="math-container">$C_{E^*}$</span> do not have a meet (= greatest lower bound) in <span class="math-container">$(\text{PropCov}(X),\leq)$</span>.<p></p> <p>Suppose that <span class="math-container">$D$</span> is a proper cover of <span class="math-container">$\omega$</span> that is a candidate for the meet of <span class="math-container">$C_E$</span> and <span class="math-container">$C_{E^*}$</span> in <span class="math-container">$(\text{PropCov}(X),\leq)$</span>. Then <span class="math-container">$D\leq C_E$</span>, <span class="math-container">$D\leq C_{E^*}$</span>, and <span class="math-container">$D$</span> is above all other common lower bounds of <span class="math-container">$C_E$</span> and <span class="math-container">$C_{E^*}$</span>, for example <span class="math-container">$C_n\leq D$</span> for all <span class="math-container">$n$</span>.</p> <p>Since <span class="math-container">$E_n=\{0,2,4,\ldots,2n\}\in C_n$</span>, there must be a set <span class="math-container">$d_n\in D$</span> such that <span class="math-container">$E_n\subseteq d_n$</span>. Since <span class="math-container">$D\leq C_E$</span>, it follows that <span class="math-container">$d_n$</span> consists of even integers. Since <span class="math-container">$D\leq C_{E^*}$</span>, it follows that <span class="math-container">$d_n$</span> is a finite set. Thus <span class="math-container">$d_n$</span> is properly contained in some <span class="math-container">$E_m=\{0,2,\ldots,2m\}$</span> for sufficiently large <span class="math-container">$m$</span>. But now repeat this argument to find some <span class="math-container">$d_m\in D$</span> containing <span class="math-container">$E_m$</span>. We now have <span class="math-container">$d_n\subsetneq E_m \subseteq d_m$</span>, contradicting the properness of <span class="math-container">$D$</span>. \\\\\</p> <p>Finally, to extend the proof that <span class="math-container">$(\text{PropCov}(\omega),\leq)$</span> is not a lattice to a proof that <span class="math-container">$(\text{PropCov}(X),\leq)$</span> is not a lattice, for any superset <span class="math-container">$X\supseteq \omega$</span>, repeat the argument above, but extend all the covers of <span class="math-container">$\omega$</span> which were used in the argument to covers of <span class="math-container">$X$</span> by adding to those covers all singletons <span class="math-container">$\{x\}$</span>, <span class="math-container">$x\in X\setminus\omega$</span>.</p> Answer 4: <p>I don't think they are isomorphic. Let $X=\mathbb N$ and let $A$ be a covering consisting of the sets of the form $$ a_n = \{ 0, \dots, n \} $$ for every $n\in \mathbb N$.</p> <p>Then, $A$ is not $\simeq$-equivalent to any proper covering.</p>
Exact bin packing the harmonic series: references?
260,525
5
https://mathoverflow.net/questions/260525/exact-bin-packing-the-harmonic-series-references
<p>Given $n$ and $B_n= \lceil H_n \rceil$, where the latter is the $n$th harmonic number $\sum^n_{i=1} 1/i$, for most $n$ it is easy to pack the first $n$ terms of the harmonic series into $B_n$ many unit bins. An interesting question is for which $n$ one cannot perform such a packing. My guess is there are no such $n$, and I would like to know of references to this problem.</p> <p>However, even more I would like to know about exact packing. There is no exact packing for $n=4$ or $5$, but (because $1= 1/2 + 1/3 + 1/6$) there are exact packings for other $n \leq 10$. Here a packing is exact if all but (at most) one bin is exactly filled to capacity.</p> <p>What is the next $n \gt 10$ for which there is an exact packing?</p> <p><strong>Edit 2017.03.09 GRP</strong>: <a href="http://oeis.org/A101877" rel="nofollow noreferrer">http://oeis.org/A101877</a> has more information. Call a subset $D$ of positive integers at most $n$ good for $n$ if the harmonic sum formed from $D$ Is floor of $H_n$. Hugo van der Sanden computed some of the good subsets below for $n=24,65,184$ and higher numbers, and Paul Hanna asked the stronger question if every subset good for $n$ had a subset good for (some number close to ) $n/e$. I am asking if there are some subsets good for enough $n$ that contain a chain of successively smaller subsets good for smaller integers so that an exact packing results. Ernie Croot III has shown (as told by Greg Martin in his ArXiv post on Denser Egyptian Fractions) a stronger result which implies that for all but finitely many integers $k$ there is an $n$ and a subset good for $n$ for which $k$ is floor of $H_n$. So far none of these references address the exact question asked above. (Double entendre intended.)</p> <p><strong>End Edit 2017.03.09 GRP.</strong></p> <p>Are there any references to this specific problem?</p> <p>Gerhard "Not Going On A Trip" Paseman, 2017.01.25.</p>
[ { "answer_id": 263363, "score": 5, "text": "<p>The next $n$ for which there is an exact packing are $24 \\leq n \\leq 30$. This is because\n$$\n1 = \\frac{1}{2} + \\frac{1}{3} + \\frac{1}{6} = \\frac{1}{4} + \\frac{1}{5} + \\frac{1}{8} + \\frac{1}{9} + \\frac{1}{10} + \\frac{1}{15}+ \\frac{1}{18} + \\frac{1}{20} + \\frac{1}{24}.\n$$\nThe sum of the remaining fractions is $&lt; 1$ for $n$ in this range. Beyond this, the problem becomes quite intractable computationally.</p>\n" }, { "answer_id": 263666, "score": 3, "text": "<p>Thanks to Jeremy Rouse, I have developed some heuristics and a nomogram-like approach to computing (bounds on a candidate for) the next smallest $n$ which admits an exact packing.</p>\n\n<p>I call them heuristics, but they are modest extensions of known facts which should be easy to prove. The primary observation is that a prime power cannot occur just once in an integral packing, and if it occurs more than once, it must occur in a cancellative combination. Thus $1/25$ or $1/27$ can't be the only denominators in an integral packing with that high a power of $5$ or $3$, nor can we just have both of $1/25$ and $1/50$, but we can have $1/27$ and $1/54$ together, and we can have $1/50$ and $1/75$ together, or we can have any of the above with enough high powers of the same prime. This follows from looking at the $p$-adic value of $a/b + c/(dp^n)$ for $c$ and $d$ coprime to $p$ and $p^n$ not dividing $b$.</p>\n\n<p>To help in determining which $n$ are feasible for an exact packing, I arrange the positive integers in several rows, with the $i$th row containing those numbers that are both multiples of the $i$th prime and have that prime as its largest prime factor. Part of the table looks like this with prime powers marked:</p>\n\n<pre><code>2 4* 8* 16*\n 3 6 9* 12 18 24 27*\n 5 10 15 20 25* 30\n 7 14 21 28\n</code></pre>\n\n<p>I then use it to find a good subset of denominators, namely a set of integers at most $n$ whose\n reciprocal sum is $B_n - 1$. The bonus is that if this subset contains another which helps form an exact packing into $B_n - 2$ bins, then this good subset helps form an exact packing for $n$.</p>\n\n<p>I can use this to find Jeremy Rouse's solution pretty quickly. As $H_{11}$ is near enough to $3 +1/11$, I first determine unfeasible $n\\gt 11$ by noting that if $n\\lt 18$, then $9$ cannot be part of a good subset,if $n \\lt 24$ this excludes $8$, and if $n\\lt 28$, then $7$ and its small multiples are excluded. Similarly, multiples of primes greater than 10 and prime powers greater than $10$ are excluded.</p>\n\n<p>This leads to the conclusion that for $n\\lt 18$ the sum of reciprocals of allowed numbers is less than $3$, and exceeds $3$ by $1/24$ only when $n=20$. When $n=24$ we get a potential good subset which excludes $12$ (because $12$ is needed to handle the excess) and Jeremy's solution pops out.</p>\n\n<p>I can use the table also to determine quickly that after $30$, the next feasible number to admit an exact packing must be greater than $51$, primarily by balancing potential members greater than $31$ of a good subset against members less than $31$ which can't be part of a good subset for $n$: if the smaller denominators overpower the larger ones, then that $n$ is not feasible.</p>\n\n<p>I then group the potential members by line, seeing if I can find subsets of each line which may be part of a good subset. Leaving out those found by Jeremy, for $n=66$ I find $16$ (contributing $1/16$), $12,27,36,48,54 (3/16)$, $30,60 (1/20),$ $7$ through $42 (7/20)$, $11,22,33,55,66 (1/5)$, and $13,26,52,65 (3/20)$, which gives an exact packing for $n=66$.</p>\n\n<p>I have not determined if $65,63,$ or $60$ are feasible, but I suspect not as multiples of $17$ and higher primes are excluded already by $n \\lt 68$, and the pair $55,66$ seems hard to replace. Note that these examples were found by hand, and that analysis of nice subsets of multiples of primes $p_i$ can be done programmatically. I disagree with Jeremy Rouse regarding the level of intractability of this problem, and thank him again for his inspiring example.</p>\n\n<p><strong>Edit 2017.03.06 GRP</strong>: After finding a packing for $n=65$ (leaving $ 21,39,44,$ and $55$ to handle the excess), I find that leaving out $65$ makes it impossible to produce an exact packing for $n=63$, as $13$ needs to be left out of the good subset, and so do a multiple of $7$ and a multiple of $11$, which is too much for the excess over $4$ of the smooth denominators that are candidates for a good packing. So after $30$, the next $n$ to admit an exact packing are $65$ through $82$.</p>\n\n<p>Based on gut feel and some trial nomogramming, I suspect the next lowest admissible $n$ will be around $170$ or greater. So far the solutions above are found by hand. The exploration continues. <strong>End Edit 2017.03.06 GRP.</strong></p>\n\n<p>Gerhard \"Who Wants To Go Further?\" Paseman, 2017.03.03.</p>\n" }, { "answer_id": 260538, "score": 0, "text": "<p>Here is the start of an approach which should lower the required ambition level of those willing to attempt a simulation.</p>\n\n<p>Consider the denominator of $d_n=H_n - B_n+1$. The definition and theory show $1 \\gt d_n \\gt 0$ for $n \\gt 1$, and so the denominator has a significant collection of prime factors. If the sum of reciprocals of those integers $i \\leq n$ which \"cover\" the prime factors of the denominator is greater than $d_n$, then $n$ cannot admit an exact cover. By this reasoning one can show there is no exact cover for 11 through 14, and many larger $n$. I can see this heading toward a proof that there are only finitely many $n$ admitting an exact packing.</p>\n\n<p>Edit: I should have thought this before posting. Apologies in advance, and thanks to any who can complete the argument.</p>\n\n<p>I believe the denominator of $d_n$ contains all but a small fraction of the primes between $n^{1/2}$ and $n$. If this is so, then the sum of the reciprocals of a set of covering integers to include in a sum for $d_n$ is (by a theorem of Mertens or Euler) is going to be not much smaller than ln 2. Throw in powers of small primes and one sees that for large $n$ admitting an exact packing, $d_n$ has to be bigger than 2/3 and likely bigger than 1. This may leave a dearth of fractions to exactly fill the next to last bin.</p>\n\n<p>Of course, the question of how many bins can be packed exactly arises. But let's wait for the first question to be resolved. End Edit.</p>\n\n<p>Gerhard \"Sorry For Not Thinking Earlier\" Paseman, 2017.01.25.</p>\n" }, { "answer_id": 263363, "score": 5, "text": "<p>The next $n$ for which there is an exact packing are $24 \\leq n \\leq 30$. This is because\n$$\n1 = \\frac{1}{2} + \\frac{1}{3} + \\frac{1}{6} = \\frac{1}{4} + \\frac{1}{5} + \\frac{1}{8} + \\frac{1}{9} + \\frac{1}{10} + \\frac{1}{15}+ \\frac{1}{18} + \\frac{1}{20} + \\frac{1}{24}.\n$$\nThe sum of the remaining fractions is $&lt; 1$ for $n$ in this range. Beyond this, the problem becomes quite intractable computationally.</p>\n" }, { "answer_id": 263666, "score": 3, "text": "<p>Thanks to Jeremy Rouse, I have developed some heuristics and a nomogram-like approach to computing (bounds on a candidate for) the next smallest $n$ which admits an exact packing.</p>\n\n<p>I call them heuristics, but they are modest extensions of known facts which should be easy to prove. The primary observation is that a prime power cannot occur just once in an integral packing, and if it occurs more than once, it must occur in a cancellative combination. Thus $1/25$ or $1/27$ can't be the only denominators in an integral packing with that high a power of $5$ or $3$, nor can we just have both of $1/25$ and $1/50$, but we can have $1/27$ and $1/54$ together, and we can have $1/50$ and $1/75$ together, or we can have any of the above with enough high powers of the same prime. This follows from looking at the $p$-adic value of $a/b + c/(dp^n)$ for $c$ and $d$ coprime to $p$ and $p^n$ not dividing $b$.</p>\n\n<p>To help in determining which $n$ are feasible for an exact packing, I arrange the positive integers in several rows, with the $i$th row containing those numbers that are both multiples of the $i$th prime and have that prime as its largest prime factor. Part of the table looks like this with prime powers marked:</p>\n\n<pre><code>2 4* 8* 16*\n 3 6 9* 12 18 24 27*\n 5 10 15 20 25* 30\n 7 14 21 28\n</code></pre>\n\n<p>I then use it to find a good subset of denominators, namely a set of integers at most $n$ whose\n reciprocal sum is $B_n - 1$. The bonus is that if this subset contains another which helps form an exact packing into $B_n - 2$ bins, then this good subset helps form an exact packing for $n$.</p>\n\n<p>I can use this to find Jeremy Rouse's solution pretty quickly. As $H_{11}$ is near enough to $3 +1/11$, I first determine unfeasible $n\\gt 11$ by noting that if $n\\lt 18$, then $9$ cannot be part of a good subset,if $n \\lt 24$ this excludes $8$, and if $n\\lt 28$, then $7$ and its small multiples are excluded. Similarly, multiples of primes greater than 10 and prime powers greater than $10$ are excluded.</p>\n\n<p>This leads to the conclusion that for $n\\lt 18$ the sum of reciprocals of allowed numbers is less than $3$, and exceeds $3$ by $1/24$ only when $n=20$. When $n=24$ we get a potential good subset which excludes $12$ (because $12$ is needed to handle the excess) and Jeremy's solution pops out.</p>\n\n<p>I can use the table also to determine quickly that after $30$, the next feasible number to admit an exact packing must be greater than $51$, primarily by balancing potential members greater than $31$ of a good subset against members less than $31$ which can't be part of a good subset for $n$: if the smaller denominators overpower the larger ones, then that $n$ is not feasible.</p>\n\n<p>I then group the potential members by line, seeing if I can find subsets of each line which may be part of a good subset. Leaving out those found by Jeremy, for $n=66$ I find $16$ (contributing $1/16$), $12,27,36,48,54 (3/16)$, $30,60 (1/20),$ $7$ through $42 (7/20)$, $11,22,33,55,66 (1/5)$, and $13,26,52,65 (3/20)$, which gives an exact packing for $n=66$.</p>\n\n<p>I have not determined if $65,63,$ or $60$ are feasible, but I suspect not as multiples of $17$ and higher primes are excluded already by $n \\lt 68$, and the pair $55,66$ seems hard to replace. Note that these examples were found by hand, and that analysis of nice subsets of multiples of primes $p_i$ can be done programmatically. I disagree with Jeremy Rouse regarding the level of intractability of this problem, and thank him again for his inspiring example.</p>\n\n<p><strong>Edit 2017.03.06 GRP</strong>: After finding a packing for $n=65$ (leaving $ 21,39,44,$ and $55$ to handle the excess), I find that leaving out $65$ makes it impossible to produce an exact packing for $n=63$, as $13$ needs to be left out of the good subset, and so do a multiple of $7$ and a multiple of $11$, which is too much for the excess over $4$ of the smooth denominators that are candidates for a good packing. So after $30$, the next $n$ to admit an exact packing are $65$ through $82$.</p>\n\n<p>Based on gut feel and some trial nomogramming, I suspect the next lowest admissible $n$ will be around $170$ or greater. So far the solutions above are found by hand. The exploration continues. <strong>End Edit 2017.03.06 GRP.</strong></p>\n\n<p>Gerhard \"Who Wants To Go Further?\" Paseman, 2017.03.03.</p>\n" }, { "answer_id": 260538, "score": 0, "text": "<p>Here is the start of an approach which should lower the required ambition level of those willing to attempt a simulation.</p>\n\n<p>Consider the denominator of $d_n=H_n - B_n+1$. The definition and theory show $1 \\gt d_n \\gt 0$ for $n \\gt 1$, and so the denominator has a significant collection of prime factors. If the sum of reciprocals of those integers $i \\leq n$ which \"cover\" the prime factors of the denominator is greater than $d_n$, then $n$ cannot admit an exact cover. By this reasoning one can show there is no exact cover for 11 through 14, and many larger $n$. I can see this heading toward a proof that there are only finitely many $n$ admitting an exact packing.</p>\n\n<p>Edit: I should have thought this before posting. Apologies in advance, and thanks to any who can complete the argument.</p>\n\n<p>I believe the denominator of $d_n$ contains all but a small fraction of the primes between $n^{1/2}$ and $n$. If this is so, then the sum of the reciprocals of a set of covering integers to include in a sum for $d_n$ is (by a theorem of Mertens or Euler) is going to be not much smaller than ln 2. Throw in powers of small primes and one sees that for large $n$ admitting an exact packing, $d_n$ has to be bigger than 2/3 and likely bigger than 1. This may leave a dearth of fractions to exactly fill the next to last bin.</p>\n\n<p>Of course, the question of how many bins can be packed exactly arises. But let's wait for the first question to be resolved. End Edit.</p>\n\n<p>Gerhard \"Sorry For Not Thinking Earlier\" Paseman, 2017.01.25.</p>\n" } ]
Question: Exact bin packing the harmonic series: references? Answer 1: <p>The next $n$ for which there is an exact packing are $24 \leq n \leq 30$. This is because $$ 1 = \frac{1}{2} + \frac{1}{3} + \frac{1}{6} = \frac{1}{4} + \frac{1}{5} + \frac{1}{8} + \frac{1}{9} + \frac{1}{10} + \frac{1}{15}+ \frac{1}{18} + \frac{1}{20} + \frac{1}{24}. $$ The sum of the remaining fractions is $&lt; 1$ for $n$ in this range. Beyond this, the problem becomes quite intractable computationally.</p> Answer 2: <p>Thanks to Jeremy Rouse, I have developed some heuristics and a nomogram-like approach to computing (bounds on a candidate for) the next smallest $n$ which admits an exact packing.</p> <p>I call them heuristics, but they are modest extensions of known facts which should be easy to prove. The primary observation is that a prime power cannot occur just once in an integral packing, and if it occurs more than once, it must occur in a cancellative combination. Thus $1/25$ or $1/27$ can't be the only denominators in an integral packing with that high a power of $5$ or $3$, nor can we just have both of $1/25$ and $1/50$, but we can have $1/27$ and $1/54$ together, and we can have $1/50$ and $1/75$ together, or we can have any of the above with enough high powers of the same prime. This follows from looking at the $p$-adic value of $a/b + c/(dp^n)$ for $c$ and $d$ coprime to $p$ and $p^n$ not dividing $b$.</p> <p>To help in determining which $n$ are feasible for an exact packing, I arrange the positive integers in several rows, with the $i$th row containing those numbers that are both multiples of the $i$th prime and have that prime as its largest prime factor. Part of the table looks like this with prime powers marked:</p> <pre><code>2 4* 8* 16* 3 6 9* 12 18 24 27* 5 10 15 20 25* 30 7 14 21 28 </code></pre> <p>I then use it to find a good subset of denominators, namely a set of integers at most $n$ whose reciprocal sum is $B_n - 1$. The bonus is that if this subset contains another which helps form an exact packing into $B_n - 2$ bins, then this good subset helps form an exact packing for $n$.</p> <p>I can use this to find Jeremy Rouse's solution pretty quickly. As $H_{11}$ is near enough to $3 +1/11$, I first determine unfeasible $n\gt 11$ by noting that if $n\lt 18$, then $9$ cannot be part of a good subset,if $n \lt 24$ this excludes $8$, and if $n\lt 28$, then $7$ and its small multiples are excluded. Similarly, multiples of primes greater than 10 and prime powers greater than $10$ are excluded.</p> <p>This leads to the conclusion that for $n\lt 18$ the sum of reciprocals of allowed numbers is less than $3$, and exceeds $3$ by $1/24$ only when $n=20$. When $n=24$ we get a potential good subset which excludes $12$ (because $12$ is needed to handle the excess) and Jeremy's solution pops out.</p> <p>I can use the table also to determine quickly that after $30$, the next feasible number to admit an exact packing must be greater than $51$, primarily by balancing potential members greater than $31$ of a good subset against members less than $31$ which can't be part of a good subset for $n$: if the smaller denominators overpower the larger ones, then that $n$ is not feasible.</p> <p>I then group the potential members by line, seeing if I can find subsets of each line which may be part of a good subset. Leaving out those found by Jeremy, for $n=66$ I find $16$ (contributing $1/16$), $12,27,36,48,54 (3/16)$, $30,60 (1/20),$ $7$ through $42 (7/20)$, $11,22,33,55,66 (1/5)$, and $13,26,52,65 (3/20)$, which gives an exact packing for $n=66$.</p> <p>I have not determined if $65,63,$ or $60$ are feasible, but I suspect not as multiples of $17$ and higher primes are excluded already by $n \lt 68$, and the pair $55,66$ seems hard to replace. Note that these examples were found by hand, and that analysis of nice subsets of multiples of primes $p_i$ can be done programmatically. I disagree with Jeremy Rouse regarding the level of intractability of this problem, and thank him again for his inspiring example.</p> <p><strong>Edit 2017.03.06 GRP</strong>: After finding a packing for $n=65$ (leaving $ 21,39,44,$ and $55$ to handle the excess), I find that leaving out $65$ makes it impossible to produce an exact packing for $n=63$, as $13$ needs to be left out of the good subset, and so do a multiple of $7$ and a multiple of $11$, which is too much for the excess over $4$ of the smooth denominators that are candidates for a good packing. So after $30$, the next $n$ to admit an exact packing are $65$ through $82$.</p> <p>Based on gut feel and some trial nomogramming, I suspect the next lowest admissible $n$ will be around $170$ or greater. So far the solutions above are found by hand. The exploration continues. <strong>End Edit 2017.03.06 GRP.</strong></p> <p>Gerhard "Who Wants To Go Further?" Paseman, 2017.03.03.</p> Answer 3: <p>Here is the start of an approach which should lower the required ambition level of those willing to attempt a simulation.</p> <p>Consider the denominator of $d_n=H_n - B_n+1$. The definition and theory show $1 \gt d_n \gt 0$ for $n \gt 1$, and so the denominator has a significant collection of prime factors. If the sum of reciprocals of those integers $i \leq n$ which "cover" the prime factors of the denominator is greater than $d_n$, then $n$ cannot admit an exact cover. By this reasoning one can show there is no exact cover for 11 through 14, and many larger $n$. I can see this heading toward a proof that there are only finitely many $n$ admitting an exact packing.</p> <p>Edit: I should have thought this before posting. Apologies in advance, and thanks to any who can complete the argument.</p> <p>I believe the denominator of $d_n$ contains all but a small fraction of the primes between $n^{1/2}$ and $n$. If this is so, then the sum of the reciprocals of a set of covering integers to include in a sum for $d_n$ is (by a theorem of Mertens or Euler) is going to be not much smaller than ln 2. Throw in powers of small primes and one sees that for large $n$ admitting an exact packing, $d_n$ has to be bigger than 2/3 and likely bigger than 1. This may leave a dearth of fractions to exactly fill the next to last bin.</p> <p>Of course, the question of how many bins can be packed exactly arises. But let's wait for the first question to be resolved. End Edit.</p> <p>Gerhard "Sorry For Not Thinking Earlier" Paseman, 2017.01.25.</p> Answer 4: <p>The next $n$ for which there is an exact packing are $24 \leq n \leq 30$. This is because $$ 1 = \frac{1}{2} + \frac{1}{3} + \frac{1}{6} = \frac{1}{4} + \frac{1}{5} + \frac{1}{8} + \frac{1}{9} + \frac{1}{10} + \frac{1}{15}+ \frac{1}{18} + \frac{1}{20} + \frac{1}{24}. $$ The sum of the remaining fractions is $&lt; 1$ for $n$ in this range. Beyond this, the problem becomes quite intractable computationally.</p> Answer 5: <p>Thanks to Jeremy Rouse, I have developed some heuristics and a nomogram-like approach to computing (bounds on a candidate for) the next smallest $n$ which admits an exact packing.</p> <p>I call them heuristics, but they are modest extensions of known facts which should be easy to prove. The primary observation is that a prime power cannot occur just once in an integral packing, and if it occurs more than once, it must occur in a cancellative combination. Thus $1/25$ or $1/27$ can't be the only denominators in an integral packing with that high a power of $5$ or $3$, nor can we just have both of $1/25$ and $1/50$, but we can have $1/27$ and $1/54$ together, and we can have $1/50$ and $1/75$ together, or we can have any of the above with enough high powers of the same prime. This follows from looking at the $p$-adic value of $a/b + c/(dp^n)$ for $c$ and $d$ coprime to $p$ and $p^n$ not dividing $b$.</p> <p>To help in determining which $n$ are feasible for an exact packing, I arrange the positive integers in several rows, with the $i$th row containing those numbers that are both multiples of the $i$th prime and have that prime as its largest prime factor. Part of the table looks like this with prime powers marked:</p> <pre><code>2 4* 8* 16* 3 6 9* 12 18 24 27* 5 10 15 20 25* 30 7 14 21 28 </code></pre> <p>I then use it to find a good subset of denominators, namely a set of integers at most $n$ whose reciprocal sum is $B_n - 1$. The bonus is that if this subset contains another which helps form an exact packing into $B_n - 2$ bins, then this good subset helps form an exact packing for $n$.</p> <p>I can use this to find Jeremy Rouse's solution pretty quickly. As $H_{11}$ is near enough to $3 +1/11$, I first determine unfeasible $n\gt 11$ by noting that if $n\lt 18$, then $9$ cannot be part of a good subset,if $n \lt 24$ this excludes $8$, and if $n\lt 28$, then $7$ and its small multiples are excluded. Similarly, multiples of primes greater than 10 and prime powers greater than $10$ are excluded.</p> <p>This leads to the conclusion that for $n\lt 18$ the sum of reciprocals of allowed numbers is less than $3$, and exceeds $3$ by $1/24$ only when $n=20$. When $n=24$ we get a potential good subset which excludes $12$ (because $12$ is needed to handle the excess) and Jeremy's solution pops out.</p> <p>I can use the table also to determine quickly that after $30$, the next feasible number to admit an exact packing must be greater than $51$, primarily by balancing potential members greater than $31$ of a good subset against members less than $31$ which can't be part of a good subset for $n$: if the smaller denominators overpower the larger ones, then that $n$ is not feasible.</p> <p>I then group the potential members by line, seeing if I can find subsets of each line which may be part of a good subset. Leaving out those found by Jeremy, for $n=66$ I find $16$ (contributing $1/16$), $12,27,36,48,54 (3/16)$, $30,60 (1/20),$ $7$ through $42 (7/20)$, $11,22,33,55,66 (1/5)$, and $13,26,52,65 (3/20)$, which gives an exact packing for $n=66$.</p> <p>I have not determined if $65,63,$ or $60$ are feasible, but I suspect not as multiples of $17$ and higher primes are excluded already by $n \lt 68$, and the pair $55,66$ seems hard to replace. Note that these examples were found by hand, and that analysis of nice subsets of multiples of primes $p_i$ can be done programmatically. I disagree with Jeremy Rouse regarding the level of intractability of this problem, and thank him again for his inspiring example.</p> <p><strong>Edit 2017.03.06 GRP</strong>: After finding a packing for $n=65$ (leaving $ 21,39,44,$ and $55$ to handle the excess), I find that leaving out $65$ makes it impossible to produce an exact packing for $n=63$, as $13$ needs to be left out of the good subset, and so do a multiple of $7$ and a multiple of $11$, which is too much for the excess over $4$ of the smooth denominators that are candidates for a good packing. So after $30$, the next $n$ to admit an exact packing are $65$ through $82$.</p> <p>Based on gut feel and some trial nomogramming, I suspect the next lowest admissible $n$ will be around $170$ or greater. So far the solutions above are found by hand. The exploration continues. <strong>End Edit 2017.03.06 GRP.</strong></p> <p>Gerhard "Who Wants To Go Further?" Paseman, 2017.03.03.</p> Answer 6: <p>Here is the start of an approach which should lower the required ambition level of those willing to attempt a simulation.</p> <p>Consider the denominator of $d_n=H_n - B_n+1$. The definition and theory show $1 \gt d_n \gt 0$ for $n \gt 1$, and so the denominator has a significant collection of prime factors. If the sum of reciprocals of those integers $i \leq n$ which "cover" the prime factors of the denominator is greater than $d_n$, then $n$ cannot admit an exact cover. By this reasoning one can show there is no exact cover for 11 through 14, and many larger $n$. I can see this heading toward a proof that there are only finitely many $n$ admitting an exact packing.</p> <p>Edit: I should have thought this before posting. Apologies in advance, and thanks to any who can complete the argument.</p> <p>I believe the denominator of $d_n$ contains all but a small fraction of the primes between $n^{1/2}$ and $n$. If this is so, then the sum of the reciprocals of a set of covering integers to include in a sum for $d_n$ is (by a theorem of Mertens or Euler) is going to be not much smaller than ln 2. Throw in powers of small primes and one sees that for large $n$ admitting an exact packing, $d_n$ has to be bigger than 2/3 and likely bigger than 1. This may leave a dearth of fractions to exactly fill the next to last bin.</p> <p>Of course, the question of how many bins can be packed exactly arises. But let's wait for the first question to be resolved. End Edit.</p> <p>Gerhard "Sorry For Not Thinking Earlier" Paseman, 2017.01.25.</p>
On the domains and extensions of unbounded operators
260,638
5
https://mathoverflow.net/questions/260638/on-the-domains-and-extensions-of-unbounded-operators
<p>I am not an expert in functional analysis but I was studying some, motivated from some mathematical physics considerations. I am not quite sure whether this is research-level, but let me state some context first:<br> An old result of functional analysis tells us that a symmetric (in the sense that $(Ax,y)=(x,Ay)$, for all $x,y \in H$), unbounded operator $A$, acting on a Hilbert space $H$, cannot be defined on the whole space but only in a dense subspace of it. This is a direct consequence of the <a href="https://en.wikipedia.org/wiki/Hellinger%E2%80%93Toeplitz_theorem" rel="nofollow noreferrer">Hellinger-Toeplitz theorem</a>. (see also: Riesz-Nagy, "Functional Analysis", 1955, p.296 and also Reed-Simon, "Methods of Modern Mathematical Physics", 1975, p.84). Since the operators of interest in physics are self-adjoint (and thus symmetric) they fall into this. </p> <p>On the other hand, it is well known that any linear map from a subspace of a Banach space $X$ to another Banach space $Y$ can be extended to a linear map $X\to Y$ defined on the whole of $X$ using Zorn's Lemma (see for example: <a href="https://mathoverflow.net/questions/31163/unbounded-linear-operator-defined-on-l2">Unbounded linear operator defined on $l^2$</a>). </p> <p><strong>Now the question is:</strong> Since the extension of a linear, unbounded operator to the whole of the space through the AC, will produce a -still- unbounded, linear operator, does the previous remark imply that the extension of linear, self-adjoint, unbounded operators on the whole of the space, produces non-self-adjoint operators? What would be a concrete relevant example? </p> <p><strong>Related question:</strong> <a href="https://mathoverflow.net/questions/178787/invertible-unbounded-linear-maps-defined-on-a-hilbert-space?rq=1">Invertible unbounded linear maps defined on a Hilbert space</a> </p>
[ { "answer_id": 260640, "score": 7, "text": "<p>Yes, you've got it right. Given an unbounded self-adjoint operator $A$ with domain $D(A) \\subset H$, using Zorn's lemma you can produce an everywhere defined operator $A'$ on $H$ which extends $A$. (In fact you can produce many such operators; the extension is highly non-unique.) By Hellinger&ndash;Toeplitz, $A'$ cannot be symmetric. So it definitely isn't self-adjoint. Another way to see this that $A'$ cannot be self-adjoint is to note that, by the closed graph theorem, $A'$ cannot be closed.</p>\n\n<p>Since you used Zorn's lemma in an essential way, you won't get a \"concrete\" description of such an $A'$. There's a strong sense in which this is true. A common working definition of \"concrete\" is \"something whose existence you can prove using only the axiom of dependent choice (DC)\". There's a famous theorem of Solovay (extended by Shelah) that it's consistent with DC that every set of reals has the property of Baire (BP); i.e. there are models of set theory in which DC and BP both hold (but full AC necessarily fails). But from BP you can prove that every everywhere defined operator on any Banach space is bounded. So in such models, $A$ won't have any extension to all of $H$. Put another way, you can't even prove $A'$ exists without using AC in an essential way, so you certainly can't construct it concretely.</p>\n\n<p>You can read more about these ideas in Schechter's <em>Handbook of Analysis and its Foundations.</em></p>\n" }, { "answer_id": 260639, "score": 6, "text": "<p>Yes, of course. By definition of the adjoint operator, $\\{[-A^* y, y]: y \\in \\mathscr D(A^*)\\}$ is the orthogonal complement in $H \\oplus H$ of the graph $\\{[x, Ax]: x \\in \\mathscr D(A)\\}$ of $A$. Thus if $A$ is unbounded and self-adjoint, i.e. $A = A^*$, the adjoint of any extension of $A$ would have a smaller domain than that of $A$. </p>\n\n<p>It's fairly easy to give concrete examples with an extension by finitely many dimensions. However, an extension to the whole space requires something like the Axiom of Choice, so you won't get a concrete example of <strong>that</strong>.</p>\n" }, { "answer_id": 260640, "score": 7, "text": "<p>Yes, you've got it right. Given an unbounded self-adjoint operator $A$ with domain $D(A) \\subset H$, using Zorn's lemma you can produce an everywhere defined operator $A'$ on $H$ which extends $A$. (In fact you can produce many such operators; the extension is highly non-unique.) By Hellinger&ndash;Toeplitz, $A'$ cannot be symmetric. So it definitely isn't self-adjoint. Another way to see this that $A'$ cannot be self-adjoint is to note that, by the closed graph theorem, $A'$ cannot be closed.</p>\n\n<p>Since you used Zorn's lemma in an essential way, you won't get a \"concrete\" description of such an $A'$. There's a strong sense in which this is true. A common working definition of \"concrete\" is \"something whose existence you can prove using only the axiom of dependent choice (DC)\". There's a famous theorem of Solovay (extended by Shelah) that it's consistent with DC that every set of reals has the property of Baire (BP); i.e. there are models of set theory in which DC and BP both hold (but full AC necessarily fails). But from BP you can prove that every everywhere defined operator on any Banach space is bounded. So in such models, $A$ won't have any extension to all of $H$. Put another way, you can't even prove $A'$ exists without using AC in an essential way, so you certainly can't construct it concretely.</p>\n\n<p>You can read more about these ideas in Schechter's <em>Handbook of Analysis and its Foundations.</em></p>\n" }, { "answer_id": 260639, "score": 6, "text": "<p>Yes, of course. By definition of the adjoint operator, $\\{[-A^* y, y]: y \\in \\mathscr D(A^*)\\}$ is the orthogonal complement in $H \\oplus H$ of the graph $\\{[x, Ax]: x \\in \\mathscr D(A)\\}$ of $A$. Thus if $A$ is unbounded and self-adjoint, i.e. $A = A^*$, the adjoint of any extension of $A$ would have a smaller domain than that of $A$. </p>\n\n<p>It's fairly easy to give concrete examples with an extension by finitely many dimensions. However, an extension to the whole space requires something like the Axiom of Choice, so you won't get a concrete example of <strong>that</strong>.</p>\n" } ]
Question: On the domains and extensions of unbounded operators Answer 1: <p>Yes, you've got it right. Given an unbounded self-adjoint operator $A$ with domain $D(A) \subset H$, using Zorn's lemma you can produce an everywhere defined operator $A'$ on $H$ which extends $A$. (In fact you can produce many such operators; the extension is highly non-unique.) By Hellinger&ndash;Toeplitz, $A'$ cannot be symmetric. So it definitely isn't self-adjoint. Another way to see this that $A'$ cannot be self-adjoint is to note that, by the closed graph theorem, $A'$ cannot be closed.</p> <p>Since you used Zorn's lemma in an essential way, you won't get a "concrete" description of such an $A'$. There's a strong sense in which this is true. A common working definition of "concrete" is "something whose existence you can prove using only the axiom of dependent choice (DC)". There's a famous theorem of Solovay (extended by Shelah) that it's consistent with DC that every set of reals has the property of Baire (BP); i.e. there are models of set theory in which DC and BP both hold (but full AC necessarily fails). But from BP you can prove that every everywhere defined operator on any Banach space is bounded. So in such models, $A$ won't have any extension to all of $H$. Put another way, you can't even prove $A'$ exists without using AC in an essential way, so you certainly can't construct it concretely.</p> <p>You can read more about these ideas in Schechter's <em>Handbook of Analysis and its Foundations.</em></p> Answer 2: <p>Yes, of course. By definition of the adjoint operator, $\{[-A^* y, y]: y \in \mathscr D(A^*)\}$ is the orthogonal complement in $H \oplus H$ of the graph $\{[x, Ax]: x \in \mathscr D(A)\}$ of $A$. Thus if $A$ is unbounded and self-adjoint, i.e. $A = A^*$, the adjoint of any extension of $A$ would have a smaller domain than that of $A$. </p> <p>It's fairly easy to give concrete examples with an extension by finitely many dimensions. However, an extension to the whole space requires something like the Axiom of Choice, so you won't get a concrete example of <strong>that</strong>.</p> Answer 3: <p>Yes, you've got it right. Given an unbounded self-adjoint operator $A$ with domain $D(A) \subset H$, using Zorn's lemma you can produce an everywhere defined operator $A'$ on $H$ which extends $A$. (In fact you can produce many such operators; the extension is highly non-unique.) By Hellinger&ndash;Toeplitz, $A'$ cannot be symmetric. So it definitely isn't self-adjoint. Another way to see this that $A'$ cannot be self-adjoint is to note that, by the closed graph theorem, $A'$ cannot be closed.</p> <p>Since you used Zorn's lemma in an essential way, you won't get a "concrete" description of such an $A'$. There's a strong sense in which this is true. A common working definition of "concrete" is "something whose existence you can prove using only the axiom of dependent choice (DC)". There's a famous theorem of Solovay (extended by Shelah) that it's consistent with DC that every set of reals has the property of Baire (BP); i.e. there are models of set theory in which DC and BP both hold (but full AC necessarily fails). But from BP you can prove that every everywhere defined operator on any Banach space is bounded. So in such models, $A$ won't have any extension to all of $H$. Put another way, you can't even prove $A'$ exists without using AC in an essential way, so you certainly can't construct it concretely.</p> <p>You can read more about these ideas in Schechter's <em>Handbook of Analysis and its Foundations.</em></p> Answer 4: <p>Yes, of course. By definition of the adjoint operator, $\{[-A^* y, y]: y \in \mathscr D(A^*)\}$ is the orthogonal complement in $H \oplus H$ of the graph $\{[x, Ax]: x \in \mathscr D(A)\}$ of $A$. Thus if $A$ is unbounded and self-adjoint, i.e. $A = A^*$, the adjoint of any extension of $A$ would have a smaller domain than that of $A$. </p> <p>It's fairly easy to give concrete examples with an extension by finitely many dimensions. However, an extension to the whole space requires something like the Axiom of Choice, so you won't get a concrete example of <strong>that</strong>.</p>
The mod p cohomologies of classifying spaces of compact Lie groups
258,497
4
https://mathoverflow.net/questions/258497/the-mod-p-cohomologies-of-classifying-spaces-of-compact-lie-groups
<p>I want to do some computation which need the mod p cohomologies of classifying spaces of connected compact Lie groups as input. I need the table for both the simply connected case and the central quotient case($G/\Gamma$, $\Gamma\subset Z(G)$). Are there a complete table for all these results. </p>
[ { "answer_id": 269980, "score": 3, "text": "<p>In the book of Mimura-Toda \"Topology of Lie groups\" you can find a lot of the computations you are interested on.</p>\n" }, { "answer_id": 271773, "score": 3, "text": "<p>Most of the time, H<sup>*</sup>(BG,𝔽<sub>p</sub>) is a polynomial algebra, generated in the degrees of the Weyl group.</p>\n\n<p>The exceptional cases are known as torsion primes. If G is simply connected and simple, there are no such primes in type A or C, they are 2 in types B (except B<sub>2</sub>), D and G, 2 and 3 in types E<sub>6</sub> and E<sub>7</sub>, and 2, 3 and 5 in E<sub>8</sub>. A precise definition is in Definition 2.43 of \"Parity Sheaves\", by Juteau, Mautner and Williamson.</p>\n\n<p>To get an idea of the cohomology in one of the unusual cases, there is\n$$H^*(BSO_3;\\mathbb{F}_2)\\cong \\mathbb{F}_2[b,w] $$\nwhere the generators b and w are in degrees 2 and 3. This is not actually hard to prove via a spectral sequence once you have the guts to believe that the cohomology ring is this big.</p>\n\n<p>For a general reference for computing this cohomology ring, look at \"Torsion in cohomology of compact Lie groups and Chow rings of reductive algebraic groups\" by Kac.</p>\n" } ]
Question: The mod p cohomologies of classifying spaces of compact Lie groups Answer 1: <p>In the book of Mimura-Toda "Topology of Lie groups" you can find a lot of the computations you are interested on.</p> Answer 2: <p>Most of the time, H<sup>*</sup>(BG,𝔽<sub>p</sub>) is a polynomial algebra, generated in the degrees of the Weyl group.</p> <p>The exceptional cases are known as torsion primes. If G is simply connected and simple, there are no such primes in type A or C, they are 2 in types B (except B<sub>2</sub>), D and G, 2 and 3 in types E<sub>6</sub> and E<sub>7</sub>, and 2, 3 and 5 in E<sub>8</sub>. A precise definition is in Definition 2.43 of "Parity Sheaves", by Juteau, Mautner and Williamson.</p> <p>To get an idea of the cohomology in one of the unusual cases, there is $$H^*(BSO_3;\mathbb{F}_2)\cong \mathbb{F}_2[b,w] $$ where the generators b and w are in degrees 2 and 3. This is not actually hard to prove via a spectral sequence once you have the guts to believe that the cohomology ring is this big.</p> <p>For a general reference for computing this cohomology ring, look at "Torsion in cohomology of compact Lie groups and Chow rings of reductive algebraic groups" by Kac.</p>
Rank of certain elliptic curves
258,606
4
https://mathoverflow.net/questions/258606/rank-of-certain-elliptic-curves
<p>I need to calculate rank of the some elliptic curves,(espicially getting generators or finding a rational point on the elliptic curves) but I cannot do this by my computer. </p> <blockquote> <p>I am interested in calculating the rank of the following elliptic curves.</p> </blockquote> <p>The elliptic curve is $$ Y^2=X^3 - (3h^2)X^2 + 3h(h^3-h)X -(h^3-h)^2 $$</p> <p>for $h=967, 1198, 1787, 1987$,</p> <p>$h=2459, 2572, 2711, 2797, 2971, 4999$</p> <p>please if possible find a generator for some of these elliptic curves.</p>
[ { "answer_id": 258698, "score": 3, "text": "<p>The following results were found using a mixture of my own codes and Denis Simon's excellent ellrank code.</p>\n\n<p>The curves have trivial torsion so Mwrank might take a long time - I haven't tried. The two heights for the rank 1 curves are for the two possible height normalisations.</p>\n\n<p>h=967 Rank=1 generator=[238501273696/245025, 900632541139856/121287375]</p>\n\n<p>h=1198 Rank=0</p>\n\n<p>h=1787 Rank=0 or 2</p>\n\n<p>h=1987 Rank=0 or 2</p>\n\n<p>h=2459 Rank=1 Height=37.4/74.8</p>\n\n<p>h=2572 Rank=2 generators= [60035809/9, 302757191/27] and [3435573760731933430513/381659437643236, 27488556048550361767336062809879/7456139229698648679016]</p>\n\n<p>h=2711 Rank =0</p>\n\n<p>h=2797 Rank=1 Height=28.1/56.2</p>\n\n<p>h=2971 Rank=0</p>\n\n<p>h=4999 Rank=1 Height=29.7/59.4</p>\n\n<p>If I get peace from my grandchildren I might try the three rank 1 curves!!</p>\n" }, { "answer_id": 258744, "score": 1, "text": "<p>The following finish off the results:</p>\n\n<p>(a) h = 2459 </p>\n\n<p>x = 2455940168334175449299068876662469864/403764781843031846693075441721</p>\n\n<p>Conductor = 329061105621720</p>\n\n<p>(b) h = 2797 </p>\n\n<p>x = 18256234369/2304</p>\n\n<p>Conductor = 550823321110248 </p>\n\n<p>(c) h = 4999</p>\n\n<p>x = 38932053386017900293094583125/1502165941669975655844</p>\n\n<p>Conductor = 401464366065</p>\n\n<p>By the way, if you set $x=z+h^2$ the curve reduces to the very simple Weierstrass form\n\\begin{equation*}\ny^2=z^3-3h^2z-h^2(h^2+1)\n\\end{equation*}</p>\n" } ]
Question: Rank of certain elliptic curves Answer 1: <p>The following results were found using a mixture of my own codes and Denis Simon's excellent ellrank code.</p> <p>The curves have trivial torsion so Mwrank might take a long time - I haven't tried. The two heights for the rank 1 curves are for the two possible height normalisations.</p> <p>h=967 Rank=1 generator=[238501273696/245025, 900632541139856/121287375]</p> <p>h=1198 Rank=0</p> <p>h=1787 Rank=0 or 2</p> <p>h=1987 Rank=0 or 2</p> <p>h=2459 Rank=1 Height=37.4/74.8</p> <p>h=2572 Rank=2 generators= [60035809/9, 302757191/27] and [3435573760731933430513/381659437643236, 27488556048550361767336062809879/7456139229698648679016]</p> <p>h=2711 Rank =0</p> <p>h=2797 Rank=1 Height=28.1/56.2</p> <p>h=2971 Rank=0</p> <p>h=4999 Rank=1 Height=29.7/59.4</p> <p>If I get peace from my grandchildren I might try the three rank 1 curves!!</p> Answer 2: <p>The following finish off the results:</p> <p>(a) h = 2459 </p> <p>x = 2455940168334175449299068876662469864/403764781843031846693075441721</p> <p>Conductor = 329061105621720</p> <p>(b) h = 2797 </p> <p>x = 18256234369/2304</p> <p>Conductor = 550823321110248 </p> <p>(c) h = 4999</p> <p>x = 38932053386017900293094583125/1502165941669975655844</p> <p>Conductor = 401464366065</p> <p>By the way, if you set $x=z+h^2$ the curve reduces to the very simple Weierstrass form \begin{equation*} y^2=z^3-3h^2z-h^2(h^2+1) \end{equation*}</p>
References studying properties of a graph which are stable under finite perturbation
258,823
4
https://mathoverflow.net/questions/258823/references-studying-properties-of-a-graph-which-are-stable-under-finite-perturba
<p>Let's say two locally finite, connected, undirected, infinite graphs are "finite perturbations" of each other if one can remove a finite subset from each and obtain isomorphic graphs (which are now possibly disconnected).</p> <p>My question is: <strong>Is there any literature on properties which are stable under this type of perturbation?</strong></p> <p>For example, does anyone know of a reference for the following statements: </p> <ol> <li>If a graph admits a non-constant bounded harmonic function then so does every finite perturbation of it.</li> <li>If the simple random walk on a graph is recurrent then so is the simple random walk on any finite perturbation.</li> <li>If the simple random walk on a graph has positive speed with positive probability, then the same holds for every finite perturbation of it.</li> <li>If the simple random walk on a graph has positive speed almost surely, then the same holds for every finite perturbation of it.</li> </ol>
[ { "answer_id": 258827, "score": 1, "text": "<p>The answers to all these questions are \"yes\" and are more or less obvious from the definitions of the corresponding properties. This is the reason why they don't appear in the literature in an explicit form. The key is property (2) - which is actually much more robust and is preserved under rough isometries (Kanai). However, the latter is not the case for the other properties.</p>\n" }, { "answer_id": 282558, "score": 1, "text": "<p>The references for these can be found in Doyle and Snell's deathless classic (which is <a href=\"https://arxiv.org/abs/math/0001057\" rel=\"nofollow noreferrer\">available for free on arXiv.org</a>. ). Section 2.4 is particularly <em>a propos.</em></p>\n\n<p><em>Doyle, Peter G.; Snell, J.Laurie</em>, Random walks and electric networks, The Carus Mathematical Monographs, 22. Washington, D. C.: The Mathematical Association of America. Distr. by John Wiley &amp; Sons, New York etc. XIII, 159 p. £ 22.00 (1984). <a href=\"https://zbmath.org/?q=an:0583.60065\" rel=\"nofollow noreferrer\">ZBL0583.60065</a>.</p>\n" } ]
Question: References studying properties of a graph which are stable under finite perturbation Answer 1: <p>The answers to all these questions are "yes" and are more or less obvious from the definitions of the corresponding properties. This is the reason why they don't appear in the literature in an explicit form. The key is property (2) - which is actually much more robust and is preserved under rough isometries (Kanai). However, the latter is not the case for the other properties.</p> Answer 2: <p>The references for these can be found in Doyle and Snell's deathless classic (which is <a href="https://arxiv.org/abs/math/0001057" rel="nofollow noreferrer">available for free on arXiv.org</a>. ). Section 2.4 is particularly <em>a propos.</em></p> <p><em>Doyle, Peter G.; Snell, J.Laurie</em>, Random walks and electric networks, The Carus Mathematical Monographs, 22. Washington, D. C.: The Mathematical Association of America. Distr. by John Wiley &amp; Sons, New York etc. XIII, 159 p. £ 22.00 (1984). <a href="https://zbmath.org/?q=an:0583.60065" rel="nofollow noreferrer">ZBL0583.60065</a>.</p>
Gaussian measure on Banach space
258,878
4
https://mathoverflow.net/questions/258878/gaussian-measure-on-banach-space
<p>Assume we have a Gaussian measure $\mu$ supported on a Banach space $X$. Can we always find a Hilbert space $H$ embedded in $X$ sch that $\mu$ is also supported on $H$?</p>
[ { "answer_id": 258892, "score": 2, "text": "<p>No, not necessarily. It is shown in Examples 3.6.6 and 3.6.7 of Bogachev's <em>Gaussian Measures</em> that if $X = C([0,1])$ and $\\mu$ is classical Wiener measure, then for any Hilbert space $H$ embedded in $X$, we have $\\mu(H) =0$, so that $\\mu$ is not supported on $H$.</p>\n" }, { "answer_id": 258956, "score": -1, "text": "<p>For the always existence of Hilbert space ‎$H‎$ need some conditions about Banach space and structure of Gaussian measure to be satisfied as montioned in the below theorem where the uniqness is the strong condition for existence but in your case it's not always Hilbert space ‎$H‎$ exist :</p>\n\n<blockquote>\n <p><strong>Theorem:</strong>\n Suppose that ‎$E‎$ is a separable, real Banach space and that is a centered Gaussian measure which is non-degenerate.Then there exists a unique Hilbert space H such that ‎$(H;E;W)‎$ is an abstract Wiener.</p>\n</blockquote>\n\n<p>For more informations look <a href=\"http://math.mit.edu/~dws/177/prob08.pdf\" rel=\"nofollow noreferrer\">this</a> paper about \"Gaussian Measures on a Banach Space\" theorem 8.2.5 with a complet proof.</p>\n\n<p><strong>Edit:</strong> I edited my answer because I don't meant the existence of Hilbert space $H‎$ related to the given question but i meant the existence will be satisfied if and only if the conditions cited in theorem 8.2.5 hold </p>\n" } ]
Question: Gaussian measure on Banach space Answer 1: <p>No, not necessarily. It is shown in Examples 3.6.6 and 3.6.7 of Bogachev's <em>Gaussian Measures</em> that if $X = C([0,1])$ and $\mu$ is classical Wiener measure, then for any Hilbert space $H$ embedded in $X$, we have $\mu(H) =0$, so that $\mu$ is not supported on $H$.</p> Answer 2: <p>For the always existence of Hilbert space ‎$H‎$ need some conditions about Banach space and structure of Gaussian measure to be satisfied as montioned in the below theorem where the uniqness is the strong condition for existence but in your case it's not always Hilbert space ‎$H‎$ exist :</p> <blockquote> <p><strong>Theorem:</strong> Suppose that ‎$E‎$ is a separable, real Banach space and that is a centered Gaussian measure which is non-degenerate.Then there exists a unique Hilbert space H such that ‎$(H;E;W)‎$ is an abstract Wiener.</p> </blockquote> <p>For more informations look <a href="http://math.mit.edu/~dws/177/prob08.pdf" rel="nofollow noreferrer">this</a> paper about "Gaussian Measures on a Banach Space" theorem 8.2.5 with a complet proof.</p> <p><strong>Edit:</strong> I edited my answer because I don't meant the existence of Hilbert space $H‎$ related to the given question but i meant the existence will be satisfied if and only if the conditions cited in theorem 8.2.5 hold </p>
Some questions about the lottery preparation
258,963
4
https://mathoverflow.net/questions/258963/some-questions-about-the-lottery-preparation
<p>I am trying to learn the lottery preparation technique from Hamkins' article but I am having some trouble understanding some of the technical details involved. I am at the beginning yet, so my difficulties are related with fast function forcing and the preservation of large cardinal notions when forcing with it. They are, probably, elemental difficulties arising from forcing but anyway I would like to ask here to get a deeper knowledge of this powerful technique.</p> <p>For example, regarding Theorem 1.4 (preservation of weakly compactness), the author asserts that ``there are only $\kappa-$many dense sets for the forcing $\mathbb{F}_{\lambda,j(\kappa)}$'' so using enough closeness property of $M$ , and $&lt;\kappa-closeness$ of $\mathbb{F}_{\lambda,j(\kappa)}$ (in fact, it is $\leq \lambda-$directed closed in $M$, though), we can get a $M-$generic $G$ in $V$. Why this quoted claim is true?</p> <p>Also, regarding Theorem 1.5 (preservation of measurability), I find a similar trouble. Hamkins claims that ``since $2^\kappa=\kappa^+$, a simple counting argument shows that $|j(\kappa^+)|^V=\kappa^+$''. Why is this true and where it is used? I suspect that it is related to dense sets and the same argument as before applies to get a $M-$generic filter. </p> <p>Every answer would be helpful. Thanks in advance!</p>
[ { "answer_id": 258990, "score": 4, "text": "<p>I assume that $j$ is the elementary map induced by a $\\kappa$-complete measure on $\\kappa$. \"$|j(\\kappa^+)|^V=\\kappa^+$\" is true because every ordinal below $j(\\kappa^+)$ can be represented in the ultraproduct by a function from $\\kappa$ into $\\kappa^+$, and the number of such functions is $(\\kappa^+)^\\kappa = \\kappa^\\kappa=2^\\kappa$. </p>\n" }, { "answer_id": 259148, "score": 1, "text": "<p>Notice that WLOG there exists at most $j(\\kappa)-$many dense sets for your forcing. Indeed, the family $D_\\alpha=\\{p:\\,\\alpha\\in dom(p)\\}$ for $\\alpha&lt;j(\\kappa)$ includes all the information for this forcing. For ''includes all the information'' I mean that $\\bigcup \\mathcal{D}=\\bigcup_{\\alpha&lt;j(\\kappa)} D_\\alpha$ (where $\\mathcal{D}$ is the family of all dense sets). Thus to get $G$ a $\\mathcal{D}-$generic filter for $\\mathbb{F}$ in $M$ it is enough to ensure that $G$ meets the collection of $D_\\alpha$'s.\nWith this at hand, regarding your first question, note that $j(\\kappa)$ is less than $\\kappa^+$ since the target model $M$ has cardinality exactly $\\kappa$. For the second one, apply the same argument and Goldstern answer and you are done.</p>\n" } ]
Question: Some questions about the lottery preparation Answer 1: <p>I assume that $j$ is the elementary map induced by a $\kappa$-complete measure on $\kappa$. "$|j(\kappa^+)|^V=\kappa^+$" is true because every ordinal below $j(\kappa^+)$ can be represented in the ultraproduct by a function from $\kappa$ into $\kappa^+$, and the number of such functions is $(\kappa^+)^\kappa = \kappa^\kappa=2^\kappa$. </p> Answer 2: <p>Notice that WLOG there exists at most $j(\kappa)-$many dense sets for your forcing. Indeed, the family $D_\alpha=\{p:\,\alpha\in dom(p)\}$ for $\alpha&lt;j(\kappa)$ includes all the information for this forcing. For ''includes all the information'' I mean that $\bigcup \mathcal{D}=\bigcup_{\alpha&lt;j(\kappa)} D_\alpha$ (where $\mathcal{D}$ is the family of all dense sets). Thus to get $G$ a $\mathcal{D}-$generic filter for $\mathbb{F}$ in $M$ it is enough to ensure that $G$ meets the collection of $D_\alpha$'s. With this at hand, regarding your first question, note that $j(\kappa)$ is less than $\kappa^+$ since the target model $M$ has cardinality exactly $\kappa$. For the second one, apply the same argument and Goldstern answer and you are done.</p>
References about tropical cluster algebras and tropical Laurent phenomenon
259,338
4
https://mathoverflow.net/questions/259338/references-about-tropical-cluster-algebras-and-tropical-laurent-phenomenon
<p>Are there some references about tropical cluster algebras and tropical Laurent phenomenon? I searched on Google but only found one paper: <a href="http://mathecon.cemi.rssi.ru/danilov/files/TP-bases.pdf" rel="nofollow noreferrer">Tropical Plucker functions and their bases </a>. </p> <p>Thank you very much.</p>
[ { "answer_id": 259353, "score": 5, "text": "<p>Let me suggest Lauren Williams and my paper as a starting point:</p>\n\n<p><em>David Speyer and Lauren Williams</em>, MR 2164397 <a href=\"http://dx.doi.org/10.1007/s10801-005-2513-3\" rel=\"noreferrer\"><strong>The tropical totally positive Grassmannian</strong></a>, <em>J. Algebraic Combin.</em> <strong>22</strong> (2005), no. 2, 189--210.</p>\n\n<p>It's earlier than a lot of the other references mentioned, so it isn't as sophisticated but, for that reason, it is a lot more concrete.</p>\n" }, { "answer_id": 259341, "score": 4, "text": "<p>There are certainly things related to tropical cluster algebras associated to surfaces. See Fomin, Shapiro, Thurston, <a href=\"https://arxiv.org/pdf/math/0608367\" rel=\"nofollow noreferrer\">https://arxiv.org/pdf/math/0608367</a>, especially section 9.5, and the sequel by Fomin and Thurston, <a href=\"https://arxiv.org/abs/1210.5569\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/1210.5569</a>, sections 14 and 15. </p>\n\n<p>There is also work by Fock and Goncharov along similar lines, see for example <a href=\"https://arxiv.org/abs/math/0510312\" rel=\"nofollow noreferrer\">https://arxiv.org/abs/math/0510312</a>. (Note that this is much more accessible than some of their work!)</p>\n" } ]
Question: References about tropical cluster algebras and tropical Laurent phenomenon Answer 1: <p>Let me suggest Lauren Williams and my paper as a starting point:</p> <p><em>David Speyer and Lauren Williams</em>, MR 2164397 <a href="http://dx.doi.org/10.1007/s10801-005-2513-3" rel="noreferrer"><strong>The tropical totally positive Grassmannian</strong></a>, <em>J. Algebraic Combin.</em> <strong>22</strong> (2005), no. 2, 189--210.</p> <p>It's earlier than a lot of the other references mentioned, so it isn't as sophisticated but, for that reason, it is a lot more concrete.</p> Answer 2: <p>There are certainly things related to tropical cluster algebras associated to surfaces. See Fomin, Shapiro, Thurston, <a href="https://arxiv.org/pdf/math/0608367" rel="nofollow noreferrer">https://arxiv.org/pdf/math/0608367</a>, especially section 9.5, and the sequel by Fomin and Thurston, <a href="https://arxiv.org/abs/1210.5569" rel="nofollow noreferrer">https://arxiv.org/abs/1210.5569</a>, sections 14 and 15. </p> <p>There is also work by Fock and Goncharov along similar lines, see for example <a href="https://arxiv.org/abs/math/0510312" rel="nofollow noreferrer">https://arxiv.org/abs/math/0510312</a>. (Note that this is much more accessible than some of their work!)</p>